content
stringlengths
1
15.9M
\section{Introduction} The electronic structure calculations in nano-scale structures have attracted considerable attention. It is much desirable to achieve molecular dynamics (MD) simulations with quantum mechanical electronic structure calculation for systems of several hundred thousands atoms with a few hundred pico-seconds (or longer time) process, since the atomic structure is very essential to the electronic structure and {\it vice versa}. Requirement for higher accuracy of the total electronic structure energy in nano-scale systems and that for larger system size contradict each other. One of possible choices for satisfying these requirements could be the LDA (local density approximation) calculation with massive-parallel computation. Even in that case, a long-time process for MD simulations is another difficulty. In the present paper, we propose one possible way to pursue a quantum mechanical MD simulation of large system size and long time process at the expense of lower accuracy of the electronic structure energy. The first-principles tight-binding method is one of promising techniques and we review the ``{\it atom-superposition and electron-delocalization molecular-orbitals theory}'' with self-consistent charge density in Section \ref{ASED}. Then, based on the tight-binding formulation, we develop novel algorithm of linear algebra for large scale systems with the non-orthogonal local orbitals in Section \ref{Krylov}. Section \ref{Application} is devoted to present applications to nano-scale systems, the fracture propagation and surface reconstruction in Si single crystals and the proton transfer in water. Section \ref{Conclusion} is conclusions. \section{First-principles tight-binding methods and charge self-consistency} \subsection{Several first-principles tight-binding methods} We have several semi-empirical tight-binding formalisms~\cite{Kwan1994,Vogl-1983} whose results are comparable to those of the first-principles LDA calculation. We have also several theories of first principles tight-binding Hamiltonian, e.g. the tight-binding LMTO method,~\cite{tb-LMTO} the LDA tight-binding method,~\cite{LDA-tb} the first principles tight-binding method of the Naval Research Laboratory (NRL).~\cite{tb-NRL} We explain in the next subsection the tight-binding Hamiltonian of the `atom-superposition and electron-delocalization molecular-orbital' theory.~\cite{ASED} \subsection{Atom-superposition and electron-delocalization molecular-orbital theory}\label{ASED} The {\it atom-superposition and electron-delocalization molecular-orbital} (ASED) theory is based on a charge-density partitioning method, where the charge density is determined by the Mulliken method of partitioning the overlap charge density (the Mulliken charge) of assumed atomic orbitals. Diagonal elements of the Hamiltonian matrix in the extended H\"uckel theory would be given as the valence orbital ionization potential with a quadratic formula of the Mulliken charge. In the present form of the ASED theory, a diagonal element of the Hamiltonian $H_{i\alpha,i\alpha}$, where $i$ and $\alpha$ denote an atom and its orbital, is set to equal to an atomic energy level (ionization energy) $\epsilon_{i\alpha}$. An off-diagonal element of the Hamiltonian is determined by the bond formation and an extended H\"uckel-like delocalization energy may be generally a good approximation;~\cite{general-Hueckel} \begin{eqnarray} H_{i\alpha,j\beta}=K(\alpha, \beta; R_{ij})\frac{\{H_{i\alpha,i\alpha}+H_{j\beta,j\beta}\}}{2}S_{i\alpha j\beta} , \label{Hap-1} \end{eqnarray} where a coefficient $K(\alpha, \beta; R_{ij})$ stands for distortion of wavefunctions depending upon orbitals and a distance $R_{ij}$,~\cite{Calzaferri-1996} and $S_{i\alpha,j\beta}$ is the overlap integral. Another contribution to the total energy is the ion-core repulsive energy, which is defined, in the tight-binding density functional (DFT) formalism~\cite{tb-DFT}, to be the difference between LDA energy and the tight-binding band structure energy. The charge transfer effects is crucial within the charge self-consistent (CSC) theory and formulated by the second order perturbation theory,~\cite{Elstner} which gives a form of the extended Hubbard-type Coulomb interaction as \begin{eqnarray} H_{i\alpha,j\beta}^{\rm CSC} = \frac{1}{2}S_{i\alpha,j\beta}\sum_{k} (\gamma_{ik}+\gamma_{jk}) \Delta q_k \label{Hap-CSC} \end{eqnarray} where $\Delta q_k$ is the deviation of the (Mulliken-type) charge of an atom $k$ and $\gamma_{ik}$ is the (diagonal or off-diagonal) Coulomb interaction between atoms $i$ and $k$ or the chemical hardness. We are applying the CSC tight-binding method to water, molecular liquids and soft materials of nano-scale size ($\simeq 3000$~atoms). A MD simulation of a proton transfer in water will be presented preliminarily in this paper. \section{Linear algebra and generalized Krylov subspace method for generalized eigen-value problem}\label{Krylov} \subsection{Linear equation and Green's function~{\rm \cite{fujiwara-2008-2009}}} In order to study electronic structures in nano-scale materials, it is much desirable to construct the order-$N$ algorithm whose computational load, e.g. cpu time and memory size, is proportional to the system size or the number of atoms. Problems on electronic properties in materials starts usually from obtaining eigen-energies and eigen-functions in materials or solving linear equations to find $| x^{(j)} \rangle$ for a given state $|j\rangle $ and a given energy $E$; \begin{eqnarray} && (E-{\hat{\cal H}}) | x^{(j)} \rangle = |j\rangle , \label{base-2} \end{eqnarray} where ${\hat{\cal H}}$ is a one-electron (Hermitian) Hamiltonian operator. Once we solve Eq.(\ref{base-2}) and find a vector $| x^{(j)} \rangle$, the matrix elements of the Green's function $ \hat{G}(z)= (z-{\hat{\cal H}})^{-1} $ can be evaluated very easily as \begin{eqnarray} && G_{ij}(E)= \langle i| (E+i\delta-{\hat{\cal H}})^{-1}|j\rangle =\langle i|x^{(j)} \rangle \label{GreensFunc} , \end{eqnarray} where $\delta$ is an infinitesimally small (positive) number. \subsection{Generalized Lanczos method~~{\rm \cite{Teng-2010}}} In case of the orthogonal basis set, the Lanczos process and the shifted conjugate orthogonal conjugate gradient (COCG) method~\cite{Vost-Melissen1990,FROMMER2003,TAKAYAMA2006} can be constructed in the framework of the Krylov subspace generated by ${\cal H}$ and ${\bm x}^{(0)}$. When the basis set $\{\phi_i\}$ is non-orthogonal, the overlap of wavefunctions $|\psi_\alpha\rangle$ ($|\psi_\alpha\rangle =\sum_i |\phi_i\rangle u_i^{(\alpha)}$) is define by the ``$S$-product'' as \begin{eqnarray} \langle \psi_\alpha | \psi_\beta \rangle = ({\bm u}^{(\alpha)}, {\bm u}^{(\beta)}) \equiv \sum_{ij} {u_i^{(\alpha)}}^*u_j^{(\beta)} S_{ij} \label{eq(0.12)} \end{eqnarray} and $S_{ij}= \langle \phi_i | \phi_j \rangle$ is an element of the overlap matrix. With a non-symmetric (non-Hermitian) matrix $H=S^{-1}{\cal H}$, we can get the three-term recursive equation or the generalized Lanczos process as \begin{eqnarray} H{\bm u}^n&=& a_n{\bm u}^n +b_{n+1}{\bm u}^{n+1} +b_n{\bm u}^{n-1} \label{eq(1.13)} \end{eqnarray} for $n=0,1,\cdots$ and $b_0=0$ with the ``S-orthogonality" as \begin{eqnarray} ({\bm u}^{n+1}, {\bm u}^{m})=0 \ \ \ {\rm for}\ \ \ \ m=n, n-1,n-2,\cdots\cdots , 0 . \label{eq(1.18)} \end{eqnarray} We can set $b_n$'s ($n=1,2,\cdots$) to be positive numbers. The generalized Krylov subspace is spanned by the vectors ${\cal H}^{m}{\bm x}^{(0)}=(S^{-1}H)^m{\bm x}^{(0)}$ and defined as \begin{eqnarray} {\cal K}_{n}({H}, {\bm x}^{(0)})={\rm span} \{{\bm x}^{(0)},{H}{\bm x}^{(0)},{H}^2{\bm x}^{(0)},\dots, {H}^{n-1}{\bm x}^{(0)}\} , \label{eq:Krylov} \end{eqnarray} Within the set of basis $\{{\bm u}^{0},{\bm u}^{1},{\bm u}^{2},\cdots\}$ or the resultant `generalized' Krylov subspace, the `Hamiltonian matrix $H$' is tri-diagonalized on the basis of $|\psi_n\rangle = \sum_i | \phi_i \rangle u_i^{(n)}$. Therefore, one can calculate the Green's function in this `generalized' Krylov subspace and also achieve the spectral decomposition (the subspace diagonalization) in the `generalized' Krylov subspace, which we call the generalized Lanczos method. \subsection{Generalized shifted COCG method~{\rm \cite{GsCOCG,Teng-2010}}} Even for non-orthogonal basis set, we can generalize the COCG procedure, similar to the original one,~\cite{TAKAYAMA2006} for the Hamiltonian ${\cal H}$ and the overlap matrix $S$. The linear equation of the `seed' energy (an arbitrary chosen energy) $\sigma_s$ can be written as \begin{eqnarray} (S^{-1}{\cal H}+\sigma_s 1){\bm x}^{(i)} = S^{-1}{\bm b} \label{shift-E} . \end{eqnarray} An approximate solution at the $n$-th iteration, a searching direction for the solution at the next iteration step, and a residual vector are written as ${\bm x}_n$, ${\bm p}_n$ and ${\bm r}_n$, respectively. We can get a recursive equations under an appropriate initial conditions; \begin{eqnarray} {\bm p}_n &=& {\bm r}_{n}^\prime + \beta_{n-1} {\bm p}_{n-1} \label{Eq:CGp:p} , \\ {\bm x}_{n+1}&=& {\bm x}_{n} + \alpha_{n} {\bm p}_{n} \label{Eq:CGp:x} , \\ {\bm r}_{n+1}&=& {\bm r}_{n} - \alpha_{n} ( {\cal H}+\sigma_sS) {\bm p}_{n} \label{Eq:CGp:r} , \end{eqnarray} with $ {\bm r}_{n+1}^\prime = S^{-1}{\bm r}_{n+1} $, $ \alpha_{n} = \frac{({\bm r}_{n}^\prime, {\bm r}_{n}^\prime)}{({\bm p}_{n}, ({H}+\sigma_s) {\bm p}_{n})} $ and $ \beta_{n} = \frac{({\bm r}_{n+1}^\prime,{\bm r}_{n+1}^\prime)}{({\bm r}_{n}^\prime,{\bm r}_{n}^\prime)}$. For non-orthogonal basis set, we can generalize the shift equation. The linear equation of a `shift' energy $\sigma$ is as follows; \begin{eqnarray} (S^{-1}{\cal H}+\sigma 1){\bm x}^{(i)} = S^{-1}{\bm b} \label{shift-E} . \end{eqnarray} Then the entire information is delivered to the coefficients $\alpha_n^\sigma$ and $\beta_n^\sigma$ at every shift energy $\sigma$ by a scalar constant $\pi^{\sigma}_{n+1}$, which is generated by the recursive equation \begin{eqnarray} \pi^{\sigma}_{n+1} &=& \{1+ \alpha_n (\sigma-\sigma_s) \} \pi^\sigma_n - \frac{\beta_{n-1}}{\alpha_{n-1}}\alpha_{n}( \pi^\sigma_{n}- \pi^\sigma_{n-1}) . \label{Eq:shift:pi} \end{eqnarray} The most important issue of the shift equation is the theorem of {\it collinear residual}:~\cite{FROMMER2003} \begin{equation} {\bm r}^\sigma_{n}=\frac{1}{\pi^{\sigma}_{n}}{\bm r}_{n} . \label{Eq:collinear} \end{equation} Then, once the solution is obtained at the seed energy by Eqs.~(\ref{Eq:CGp:p})$\sim$(\ref{Eq:CGp:r}), solutions at arbitrary shift energies can be evaluated by using only scalar-scaler and scalar-vector multiplications. The only remaining problem is the calculation of $S^{-1}{\bm r}_n$. Fortunately, the overlap matrix $S$ is positive-definite and nearly equal to the unit matrix. Therefore, we can solve this equation efficiently by the CG method, e.g. solve inversely ${\bm r}_n =S{\bm r}_n^\prime$ for a given ${\bm r}_n$. The choice of the seed energy $\sigma_s$ is not unique. If one would choose a seed energy in an energy range of faster convergence, spectra at majority energy points have not be converged yet after achieving the convergence at the seed energy. In that case, one might restart the calculation with new seed energy from the beginning. The {\it seed-switching} is very efficient technique to avoid waste of the already finished part of constructing the Krylov subspace.~\cite{SOGABE2007,yamamoto2007b} One can choose a new seed energy $\sigma_s^{\rm new}$ and can continue the calculation without discarding the information of the previous calculation with the old seed energy $\sigma_s$. \begin{figure}[htbp] \begin{center} \resizebox{0.7\textwidth}{!}{ \includegraphics{d-otbital-lt.eps} } \end{center} \caption{\label{FIG-lDOS-test} The local density of states of fcc gold of Au~256 atoms by the tb-LDA with non-orthogonal basis set. There are three lines in the figure and they are almost overlapped. The chemical potential locates at $+0.2947$~Ry. } \end{figure} \subsection{Example} We have explained the generalized Lanczos method and the generalized shifted COCG method. Now we present the calculated results of the local density of states (LDOS) by using the tight-binding Hamiltonian with non-orthogonalized basis set given by the NRL Hamiltonian.~\cite{tb-NRL} The system is of gold 256 atoms in a supercell of an fcc structure with s-, p-, and d-orbitals under a periodic boundary condition. Figure~\ref{FIG-lDOS-test} shows the LDOS of d-orbitals, where the vertical axis is in logarithmic scale to see the behavior of tails. The three calculated results, the generalized Lanczos method, the generalized shifted COCG method and the exact calculation, are almost identical with each other. The cpu times by a standard work-station of a single processor are 3.92~s, 21.20~s and 57.6~s for the generalized Lanczos method, the generalized shifted COCG and the exact calculation, respectively. The algorithm of the present developed algebraic methods is suitable for parallel computation. on the other hands, the cpu time grows as a cubic power of the matrix size in the exact calculation. Therefore, the present methods are promising in large scale electronic structure calculations. \section{Application to nano-scale systems}\label{Application} \subsection{Scale-size dependence of recently developed linear algebraic methods} Much attention has been paid to nano-scale systems and the first principles molecular dynamics simulation has become more important in material sciences and engineering. We have developed a set of computational methods of electronic structure calculations, {\it i.e.} the generalized Wannier state methods,~\cite{HOSHI2000,HOSHI2003,HOSHI2005,HOSHI2006} the Krylov subspace method,~\cite{TAKAYAMA2004} and the shifted COCG method for nano-scale systems.~\cite{TAKAYAMA2006} In the preceding sections, we have explained recent development on the ASED and generalized shifted COCG method. The tight-binding MD simulation method with being-developed linear algebraic algorithm is useful to achieve calculations in extra-large systems whose computational cost increases linearly proportional to the system size as shown in Fig.\ref{FIG-BENCH}, where tested examples are metallic, insulating and semiconducting dense materials. \begin{figure}[thbp] \begin{center} \resizebox{0.48\textwidth}{!}{ \includegraphics{bench-2005-10-05-mono-lt.eps} } \end{center} \caption{\label{FIG-BENCH} The computational time as a function of the number of atoms ($N$). \cite{HOSHI2003, HOSHI2005, HOSHI2006} The time was measured in a older standard workstation for metallic (fcc Cu and liquid C) and insulating (bulk Si) systems with up to 11,315,021 atoms, by the conventional eigenstate calculation (EIG) and by our methods for large systems; KR-SD (Krylov subspace-diagonalization), WS-VR (Wannier-state variational) and WS-PT (Wannier-state perturbation) methods. The CPU time increases linearly with increasing the number of atoms. See the original papers~\cite{HOSHI2003, HOSHI2005, HOSHI2006} for the details of parallel computation. } \end{figure} \subsection{Fracture propagation and surface reconstruction in silicon crystal}\label{Si} Here, we will show a calculation results of fracture propagation in a bulk Si single crystal with the semi-empirical tight-binding Hamiltonian by Kwon {\it et al}~\cite{Kwan1994}, based on the orthogonal basis set. The cleavage propagation speed is estimated, in the present model, as $v_{\rm prop} \simeq 2~{\rm nm/ps} = 2~{\rm km/s}$, which is comparable to the velocity of the Rayleigh wave $v_{\rm R}=4.5~{\rm km/s}\simeq 4.5~{\rm nm/ps}$.~\cite{HOSHI2003,HOSHI2005} The easy-propagating plane of fracture in Si is known widely to be that of (110) or (111) and we studied the phenomena of fracture propagation in 14~nm scale Si crystals.~\cite{HOSHI2005} In case of fracture on the (111) plane, the the Pandy structure~\cite{PANDEY} of (111)-$(2\times 1)$ surface reconstruction appears with several steps. The direction or the structure of fracture propagation planes is not explained by the stabilization energy of the final stable surfaces structure because a transient surface structure is first formed immediately after the bond-breaking and, after appearance of the ideal surface, the surface reconstruction happens to form the final reconstructed surface. In the process of fracture propagation and surface reconstruction, the two different kinds of energy loss and gain compete with each other, e.g. the energy competition between an electronic energy gain of breaking bonds and an elastic energy loss of distorting the lattice. This kind competition may be possible only in nano-scale systems and, in a MD simulation, we should prepare larger systems, e.g. a few tens nm length. Then, even a fracture propagation starts on a (001) plane, the plane of the fracture propagation changes to (111) and (110) planes. Figure \ref{Frac-Prop-111} shows examples of the surface reconstruction after formation of surfaces. \begin{figure}[t] \begin{center} \includegraphics[width=0.9\linewidth]{fracture-prop-lt.eps} \end{center} \caption{ Fracture propagation on a (111) plane of a Si single crystal. (a) Appearance of Pandy structure without steps. (b) Pandy structure with a step. The numbers show the number of atoms in a ring. The symbols ($+$) or ($-$) refer to the electron charge, excess or deficiency.} \label{Frac-Prop-111} \end{figure} \subsection{Proton transfer in water}\label{H3O-nH2O} The correlation energy of water dimer,~\cite{MCC-theory} hydrogen-bonded network~\cite{hydrogen-bond} and protonated water~\cite{protonated-water} are all long-investigated problems of quantum chemistry. We have been applying the ASED method to water and proton transfer in water to investigate the applicability of the method. The bond angle H-O-H (an experimental value 104.5$^\circ$) is determined sensitively by the energy levels (binding energy and bond angle dependence) 1B$_2$ and 3A$_1$ of H$_2$O, which could be done by adjusting the ionic energy levels $\epsilon_{\rm 2s}$ and $\epsilon_{\rm 2p}$ of oxygen 2s and 2p orbitals and their STO parameters. After that, we could obtain the correct order of molecular energy levels of (2A$_1$), (1B$_2$), (3A$_1$), (1B$_1$), (4A$_1$), (2B$_2$) and an approximate cohesive energy of a molecule (10.97~eV, Cf. the experimental value of 10.08~eV). The bond length can be adjusted by slight rescaling the two-body repulsive term and the resultant bond length is 3.168\AA and the CSC scheme works efficiently (Cf. 2.797\AA without CSC scheme and 2.967\AA by Gaussian B3LYP/6-31G(d,p)). Once we include H 2p state in the calculation, the molecular dimer configuration $\alpha=7^\circ$, $\beta=57^\circ$ can be obtained, which is in good agreement with the results of Gaussian calculation ($\alpha=6^\circ$ and $\beta=57^\circ$. See the inset in Fig.~\ref{Proton}). \begin{figure}[t] \begin{center} \includegraphics[width=0.9\linewidth]{Proton-lt.eps} \end{center} \caption{ Snap shot of the proton transfer process. A proton transfers from a H$_2$O molecule A to that of B and another proton will transfer from the oxonium ion H$_3$O$^+$ to that of C. The inset shows the definition of angles $\alpha$ and $\beta$ for a stable configuration of a water dimer.} \label{Proton} \end{figure} Then we simulate the proton transfer in a water system 100H$_2$O + H$_3$O$^+$ and observe the proton transfer process to happen in a recombination of H$_3$O$^{+}$ + H$_2$O $\Rightarrow$ H$_2$O + H$_3$O$^{+}$ (hopping of a proton) but not the transfer of an oxonium ion H$_3$O$^{+}$ itself and the jumping time is of the order of 1~ps. Furthermore, we observe the following elementary process within an order of 0.1~ps;\\ (1) A H$_3$O$^{+}$ ion tries to find a proper one in near-neighboring H$_2$O molecules, \\ (2) a H$_3$O$^{+}$ ion forms a weak bond with one of H$_2$O (forming H$_3$O$^+$+H$_2$O), and \\ (3) a proton transfers to that H$_2$O, forming H$_2$O+H$_3$O$^+$. \\ (4) If this H$_2$O is proper one, the new H$_2$O molecule (the former H$_3$O$^{+}$ ion) rotates (to form a stable water dimer 2H$_2$O) and change to a stable configuration and, at the same time, \\ (5) a new H$_3$O$^{+}$ ejects a proton H$^+$ to a next H$_2$O. From these observation, we may conclude that the ASED framework can work properly in a wide variety of materials. The present system size of water is still too small to use the generalized Lanczos method and the generalized shifted COCG method. Even so, we have already assured these methods for working in the MD simulation in the present system. \section{Conclusions}\label{Conclusion} We have reviewed our recently developed methods of ASED based on the generalized H\"uckel approximation and the novel linear algebraic algorithm for large-scale quantum molecular dynamics simulation. The crucial point is that the novel linear algebraic algorithm is applicable to systems of several thousand atoms or more with the same accuracy of the exact calculation. Then we presented examples of the applications to the fracture propagation and surface reconstruction in Si and the proton transfer in water. The water is very sensitive and difficult systems to simulate and the successful results suggest that our proposed simulation method is promising in a wide variety of systems, dense or dilute, solids, liquids or soft materials, and metals or insulators. The present MD simulation method was applied also to the problems of the formation of helical multishell Au nanowire~\cite{Au-nanowire} and the diamond-graphite transformation under applied stress~\cite{diamond-graphite}. \section*{Acknowledgments} Numerical calculation was partly carried out using the supercomputer facilities of the Institute for Solid State Physics, University of Tokyo and the Research Center for Computational Science, Okazaki. Our research progress of large-scale systems and other information can be found on the WEB page of ELSES (Extra-Large Scale Electronic Structure calculation) Consortium http://www.elses.jp . \section*{References}
\section{Introduction} Problems of packing and space tiling have fascinated scientists for a very long time. Kepler's 1611 essay {\it On the Six-cornered Snowflakes} is probably one of the earliest publications on the subject. Here he conjectured that cubic close packing and hexagonal close packing are the most efficient ways to fill a space using equally sized spheres. It wasn't until 1998 that Kepler's conjecture was finally announce to be proven (with 99\% degree of confidence) by Thomas Hales~\cite{Hales}. Recent advances in particle synthesis~\cite{DeVries,Schnablegger,Hong,Weller,Hobbie} have opened the way to the production of colloidal particles that are anisotropic both in shape and surface chemistry. This provides an unlimited number of building blocks that can spontaneously organize into an unprecedented variety of structures with potentially novel functional, mechanical, and optical properties. For these reasons, the problem of efficient packing of nanoparticles is under intense investigation. Most of the work on particle crystallization and self-assembly in the last decade has focused on monodisperse~\cite{frenkel1} or polydisperse~\cite{frenkel2,zaccarelli} systems of spherical or regularly shaped particles (see also~\cite{glotzer,chandler,geissler,cacciuto,torquato,frenkel,glotzer2,glotzer3,glotzer4,esco,weitz} and references therein). Nevertheless, there are several important cases in which the shape of the single components cannot be tailored at will; however, an efficient packing, or an understanding of the physical properties of these densely compressed systems, is highly desirable. Two examples of outstanding problems in this category are the storage of grains~\cite{deGennes} and protein crystallization~\cite{rosenberger}. Both examples can be ideally thought of as two different aspects of the problem of understanding the role of shape in particle packing. In the first case the goal is to efficiently pack a system of randomly-shaped polydisperse grains. In the second case the aim is to crystallize non-spherical, yet equally shaped monodisperse components. The latter case is the focus of the present paper. We want to understand how the ability of particles to form macroscopically ordered crystal structures is affected by their shape. Specifically, we analyze how random perturbations from the ideal spherical shape affect the crystallizability of a densely packed system of indistinguishable hard particles. Although a few papers have dealt with the thermodynamic behavior of soft/deformable particles as a model for polymer brushes or polymer-coated colloids (\cite{pamies,capone,richter,bozorgui,ziherl} and references therein), to the best of our knowledge, this paper is the first study where shape distortions, frozen onto the particles, are explicitly and systematically accounted for. The problem of crystal formation is poorly understood. From a thermodynamic standpoint, classical nucleation theory informs us that a crystal can only be formed via a barrier crossing event. The free energy gain to form a nucleus of a stable crystalline structure in a supersaturated solution has to balance out the free energy cost associated with the formation of an interface between the solid and the fluid parent phase. The Gibbs free energy cost, $\Delta G$, associated with this event has a strong dependence on the interfacial free energy, $\gamma$: ${\Delta}G \propto \frac{\gamma^3}{(\rho\Delta\mu)^2}$, where $\rho$ and $\Delta\mu$ are the density and the chemical potential difference between solid and fluid phase, respectively. As $\gamma$ is very hard to extract from experiments, computer simulations are the tool of choice to study and analyze this important process in detail. One way to investigate our problem could be to measure the dependence of $\gamma$ (or $\Delta G$) on the magnitude of the shape perturbations at constant $\Delta\mu$. This approach would be appropriate when considering a system that is polydisperse in shape and size; however, in our case, we have to study a statistically relevant number of systems each containing $N$ identical particles with the same shape perturbation. Such an approach is therefore impractical because (a) it would require the computation of a phase diagram for every system with a given particle realization (about 500 in this study) and (b) for sufficiently large shape deformations the FCC structure that nucleates out of a solution of hard spheres may not necessarily be the most stable one, and this would lead to a meaningless comparison of unrelated $\gamma$s. Given these limitations, our strategy will be to generate a large number of particle shapes, sort them in terms of the extent of their asphericity, and simply analyze whether they organize into an ordered and periodic structure (whatever that might be) once compressed at large densities. We are left with the problem of devising a new model that allows for an effective control of the particle's degree of asphericity. Unfortunately there is not a unique way of introducing perturbations on the particle shape, so we have opted for a model that, without loss of generality, weds simplicity and numerical efficiency. \section{Model} In our model, each particle is built by setting the center of $N_b$ ($4 \leq N_b \leq 12$) spheres of diameter $\sigma$ {at random positions} inside a spherical shell of diameter $\sigma_0<\sigma$. The overall volume generated from the resulting overlapping aggregate defines our new particle. Deviations from the ideal spherical shape can be conveniently controlled by varying $\sigma_0$ and $N_b$. {Values of $\sigma_0$ and $N_b$ were chosen to achieve wide coverage of the range of possible particle geometries.} For $\sigma_0=0$ one recovers the spherical limit, and as $\sigma_0$ increases, particles develop larger and larger shape distortions. In a similar fashion, large values of $N_b$ result in a bumpy but overall isotropic particle, whereas small values of $N_b$ tend to generate very anisotropic shapes. Once a particle is built, the center of mass of this cluster of balls is determined, and the entire cluster is scaled so that its total volume equals that of a spherical particle of diameter $\sigma$, i.e. $\frac{\pi}{6}\sigma^3$. Any two particles $i$ and $j$ interact via a hard repulsive potential defined as \begin{equation} U_{ij}= \begin{cases} 0 & {\textrm{if }} |r_s-r_t|>\sigma_R \,\,\,\,\,\,\forall s\in i \,\,,\,\, \forall t\in j \cr \infty &\text{otherwise}\cr \end{cases} \end{equation} where $s$ and $t$ run over all spheres of rescaled diameter $\sigma_R$ constituting particle $i$ and particle $j$ respectively. Experimental realizations of colloidal particles similar to ours could be generated using the approach described in reference~\cite{weitz} to create uniform nonspherical particles with tunable shapes. Figure 1 shows a few snapshots of particle shapes obtained for different values of $N_b$ and $\sigma_0$. \begin{figure} \includegraphics[width=0.36\textwidth]{parts.eps} \caption{Model particles for different values of $\sigma_0$ and $N_b$ built according to the scheme described in the text.} \end{figure} To analyze the high density behavior of these systems we used Monte Carlo simulations in the $NPT$ ensemble. Given the large number of simulations involved in this study, we limited the size of our systems to a total of $N=128$ particles. {We use a cubic box with periodic boundary conditions.} Using as a reference the hard-spheres model, i.e. coexistence reduced pressure and volume fraction respectively at $P^*\simeq 11.6$ and $\phi\simeq 0.494$, we performed, for each system with a given particle realization, a series of simulations at increasing values of pressure beginning at $P^*=10$ with a typical increment of $\Delta P^*$=0.5 or smaller. We ran each simulation for a minimum of $4\cdot 10^6$ Monte Carlo sweeps after thermalization. The largest pressure was set by either the crystallization of the system, or by the reaching of a volume fraction $\phi\simeq 0.6$, above the glass transition point of hard spheres $\phi_G\simeq 0.58$. Crystallization was detected in our samples with a combination of methods: (a) standard spherical-harmonics based bond order parameter $q_6$~\cite{q61,q62}, (b) a careful monitoring of the system volume fraction over time for sudden jumps, and (c) visual inspection. We investigated a total of 487 different particle shapes. We find that at least two order parameters are required to properly characterize the shape of each particle. The first is its asphericity $A$ (as opposed to the commonly used sphericity, $S$~\cite{sphericity}), defined in terms of the surface to volume ratio of a particle $\alpha_p=A_p/V_p$ with respect to that of a sphere of diameter $\sigma$, $\alpha_s=6/\sigma$, as \[A = 1 - S = 1-\frac{\alpha_s}{\alpha_p}.\] Given our model setup, $V_p=V_s$, $A$ simplifies to $A=1-\pi\sigma^2/A_p$ The second parameter, $q$, measures the orientational symmetry of the particle. It is used to describe the asphericity of random walks~\cite{rudnick}, and it is obtained by combining invariants of the particle inertia tensor $I_{ij}$ as \[q=\frac{\left(R_1^2-R_2^2\right)^2 + \left(R_1^2-R_3^2\right)^2 + \left(R_2^2-R_3^2\right)^2}{2\left(R_1^2 + R_2^2 + R_3^2\right)},\] where $R_1$, $R_2$, and $R_3$ are the three principal eigenvalues of the inertia tensor of the particle, i.e., the three principle radii of gyration of the particle. Both parameters are defined so that they are equal to $0$ for a perfectly spherical particle, and approach $1$ for extremely aspherical ones. Note that $A$ depends intimately on the value of $\sigma_0$ used to construct the particle, whereas $q$ depends only on the angular distribution of spheres about the center of mass -- that is, it is completely independent of $\sigma_0$. \section{Results} Figure~\ref{plot} summarizes the main result of our paper. \begin{figure} \includegraphics[width=0.5\textwidth]{data.eps} \caption{Crystallizability of aspherical particles characterized in terms of two shape parameters $A$ and $q$. Filled circles indicate particles that {easily} crystallized, while open squares indicate particles that did not. {Data include results for 487 total particle geometries, including 85 using $N_b=4$, 200 using $N_b=6$, 92 using $N_b=8$, and 110 using $N_b=12$.} The solid line is a guide to the eye.}\label{plot} \end{figure} It is obtained by collecting the crystallizability of the 487 systems built out of the 487 different particle geometries we have generated across the $A$, $q$ spectrum. Each point in the $A$ vs $q$ diagram represents the result of a set of simulations at different pressures. As would be expected, crystallization is favored when both $A$ and $q$ are small $-$ that is, when the particles are nearly spherical by both measures. A roughly inverse relationship is clearly evident; particles with large $A$ must have very small $q$ in order to have a hope of crystallization, and vice-versa. But more importantly, we find the existence of a clear boundary delineating the crystallizability limit for every possible shape generated with our model (small deviations at the interface are likely due to finite size effects and or the limited length of our simulations). This is quite remarkable because it provides a very useful way of predicting whether a particular particle shape can pack into a crystalline structure by simply measuring the experimentally accessible $A$ and $q$. We find that a good empirical expression to describe the phase boundary is $A(q) = 0.023 + 1/(170q-10)$. This curve is only meant to serve as a guide to the eye and to give an approximate numerical estimate of the location of the phase boundary for $q\gtrsim 0.08$. Notice that for smaller values of $q$ our data seem to indicate a sharp end of the crystal boundary. We believe this to be an artifact of our particle model. In fact, that region is where the value of $\sigma_0$ becomes sufficiently large ($\sigma_0>0.5\sigma$) to break the compactness of the particles generated with our method, especially those with smaller values of $N_b$. Furthermore, as small values of $q$ indicate a large orientational symmetry, we expect this region to be heavily populated by specific geometric arrangements, whose packing properties, at large values of $A$, will be extremely sensitive of the particular value of $N_b$. As our model is intended to describe randomly shaped particles, our diagram does not include the results for particles designed with very specific shapes such as rods, plates or regular polyhedric geometries that are known to crystallize. These particular cases would generate sharp peaks around specific values of $q$. Furthermore, is not clear that our two order parameters, that have after all been selected to describe asphericity, would be the most appropriate to study deviations from an arbitrary non spherical designed shape. We therefore limited our study to $N_b >3$, to explicitly avoid trivial cases as rod-like ($N_b=2$) and plate-like particles ($N_b=3$). Two obvious questions present themselves in the face of these data. The first is: what is the nature of the crystals formed when particles do {easily} crystallize; does the system present translational but not orientational order, as expected for $A\simeq 0$, or does the rotational motion of the particles becomes restricted for large values of $A$? The second is: what sets the boundary between the two phases; do the particles that fail to {easily} crystallize do so because they become kinetically trapped or because of the lack of a stable crystal phase? \begin{figure} \subfigure{ \includegraphics[width=0.4\textwidth]{rot.eps} \put(-200,140){\bf{(a)}} \label{rot} } \subfigure{ \includegraphics[width=0.4\textwidth]{trans.eps} \put(-200,140){\bf{(b)}} \label{trans} } \caption{(a) Rotational mean square displacement for a subset of the particles considered. The solid line is a reference for nearly spherical particles. The dashed line is the result for those particles that {easily} crystallized and the dotted line shows $\left<\Delta\theta^2\right>$ for those that did not. The data were averaged over 20 realizations in each region under the same pressure $P^*=20$. (b) Translational mean square displacement for 20 different particle shapes that failed to {easily} crystallize.}\label{diff} \end{figure} To address the first question, we measured the rotational diffusion, $\left<\Delta\theta^2\right>$, for 20 systems that {easily} crystallize, all at a reduced pressure $P^* = 20$ and near the phase boundary. As a reference we also plotted $\left<\Delta\theta^2\right>$ for nearly spherical particles, obtained by setting $\sigma_0 = 10^{-4}\sigma$, where we know particles are free to rotate at their lattice sites. As can be seen in Fig.~\ref{rot}, we find no evidence that particles in the crystalline phase become orientationally arrested or manifest an orientationally anomalous behavior. The only effect is that of decreasing their diffusion constant, but this is expected from simple geometrical considerations. It is obvious that at very large densities, regardless of the specific phase a system selects, particles' orientations will manifest a glassy behavior or eventually freeze.~\cite{schweitzer} It is therefore of interest to also look at the dynamical properties of those particles in systems do not {easily} crystallize and are located just across the phase boundary from the ones that do {easily} crystallize. These results, obtained at the same reduced pressure $P^*=20$, are also shown in Fig.~\ref{rot}. We find no signature of anomalous dynamics, neither in the rotational (Fig.~\ref{rot}), nor in the translational (Fig.~\ref{trans}) degrees of freedom, i.e. such systems behave as regular fluids. This result is by no means conclusive, as a thorough investigation of this last point would require larger system sizes and an event-driven dynamics of the components. Our results seem to suggest that what sets the location of the phase boundary is not a sudden slowdown of the dynamics of the system, but more likely an increase of the Gibbs free energy difference between the crystalline and the fluid phase, analogously to that found for polydisperse spherical particles~\cite{frenkel2}. It would be interesting to investigate the equilibrium properties and the stability of candidate crystalline structures for different values of $A$ and $q$ to investigate whether our boundary line coincides with the onset of crystal instability, but for the reasons given above, we have not attempted to do so. \section{Conclusion} Apart from the details concerning the dynamics of these systems, our results show that, for aspherical particles, shape and crystallizability can be directly and easily correlated when particles are characterized in terms of their asphericity via $q$ and $A$. {It is possible that for some highly asymmetric particle geometries a non-cubic box would accommodate crystallization whereas a cubic one would not; however, we believe this is a secondary effect and its relevance is likely within the range of errors due to finite size effects.} Our data suggest precise limits for the manufacture of nanocomponents expected to crystallize, and may have important implications for the problem {of} protein crystallization. The latter system is clearly far more complex than the one we explored, since not only shape, but also interparticle direct interactions (which are not necessarily isotropic), are responsable for the organization of proteins into large macroscopic crystals. Nevertheless, it would be interesting to systematically explore to what extent a similar correlation exists in this case. \section*{Acknowledgments} This work was supported by the National Science Foundation under CAREER Grant No. DMR-0846426.
\section{Introduction} \section{The modified MA model and its approximations} Space-like observables $\mathcal{D}(Q^2)$ are analytic in the entire complex $Q^2$-plane with the exception of the negative semiaxis starting at some threshold $-M^2_0<0$. In standard perturbative QCD (pQCD), the coupling $a(Q^2) \equiv \alpha_s(Q^2)/\pi$ has Landau singularities at low energies $Q^2 \sim \Lambda^2 > 0$. This nonanalytic behavior is reflected in the pQCD-evaluated expressions of the aforementioned space-like observables $\mathcal{D}(Q^2)$, contravening the analyticity properties of the true $\mathcal{D}(Q^2)$. The aim of analytic QCD (anQCD) models is to provide a coupling $A_{1}^{(an)}(Q^2)$ with the correct analytic structure. We will consider the following expression for the QCD running coupling: \begin{equation} \label{mMA} A_{1}^{(an)}(Q^2)= \frac{1}{\pi} \int_{M_{0}^{2}}^{\infty}d\sigma\frac{\rho_{1}(\sigma)}{\sigma+Q^2}, \end{equation} where $\rho_1(\sigma)={\rm Im}[a(-\sigma-i\epsilon)]$ and $a$ is the pQCD running coupling. In the pQCD case $M_0^2=-\Lambda^2<0$. In the widely used Minimal Analytic (MA) model \cite{Shirkov} $M_0^2=0$. The MA coupling is analytic down to $Q^2=0$; however, the latter point is not in the analyticity domain because the derivatives there diverge. We consider a modification of MA, $A_1^{(mMA)}$ with $M_0^2>0$, as introduced in \cite{Nesterenko:2004tg}. This coupling has the same type of analyticity domain as the space-like observables. For the purpose of integration, $\rho_1(\sigma)$ can be well approximated as a sum of positively weighted Dirac deltas (for application of such approximations in a somewhat different context see Ref.\cite{Peris:2006ds}). This approximation is equivalent to the paradiagonal Pad\'e approximant $R_{M}^{M-1}(Q^2)$ of $A_1^{(mMA)}$ \cite{Cvetic:2009mq} \begin{equation} \label{narrow} \frac{\rho_1(\sigma)}{\pi} \ \Theta(\sigma-M_0^2) \approx \sum_{n=1}^{M}F_{n}^{2}\delta(\sigma-M_{n}^{2}) \ \Rightarrow \ A_{1}^{(mMA)}(Q^2)\approx \sum_{n=1}^{M}\frac{F_{n}^{2}}{M_{n}^{2}+Q^2}. \end{equation} We can use these approximants to evaluate the coupling and avoid the more time-consuming calculation of the dispersive integral in Eq. (\ref{mMA}). Further, mMA coupling is a Stieltjes function, so all the poles $Q^2_{\rm pole}$ of its paradiagonal Pad\'e approximants lie on the negative real axis ($Q^2_{\rm pole} < -M_0^2$) keeping the desired analytic structure. \begin{figure}[htb] \begin{minipage}[b]{.49\linewidth} \centering\includegraphics[width=1\textwidth]{pade} \end{minipage} \begin{minipage}[b]{.49\linewidth} \centering\includegraphics[width=1\textwidth]{dmMA} \end{minipage} \vspace{-0.2cm} \caption{ Left: the full line is the mMA coupling, the dashed lines are the Pad\'e approximants $R^{M-1}_{M}$: $R^{0}_{1}$, $R^{1}_{2}$, ..., $R^{9}_{10}$, with $M_0=2M_{\pi}$; right: pQCD coupling full line, MA dotted, dmMA dashed.} \label{figure01} \end{figure} \section{The delta-modified MA model} Motivated by the success of the approximation of $\rho_1$ by a sum of Dirac deltas, and with the aim of better reproducing data, an additional Dirac delta in Eq.~(\ref{narrow}) can be introduced to parametrize the unknown physics in the domain of low positive $\sigma$: $\Delta \rho_1(\sigma) = \pi F_{-1}^2 \delta(\sigma-M_{-1}^2)$. Here, $F_{-1}^2$ and $M_{-1}^2$ are new positive parameters ($0< M_{-1}^2 < M_0^2$) of this ``delta-modified'' MA model (dmMA). This leads to the coupling \begin{equation} \label{dmMA} A_1^{(dmMA)}(Q^2)=\frac{F_{-1}^2}{Q^2+M_{-1}^2}+ \frac{1}{\pi} \int_{M_0^2}^{\infty}d \sigma \frac{\rho_1(\sigma)}{\sigma+Q^2}. \end{equation} The parameters $F_{-1}^2$ and $M_{-1}^2$ can, for example, be fixed by requiring that $|A_1^{(dmMA)}(Q^2)-a(Q^2)|\sim (\Lambda^2/Q^2)^3$ for $Q^2\rightarrow\infty$, i.e., dmMA being much closer to pQCD in the asymptotic region than the MA and mMA couplings ($\Rightarrow$ the value of $\Lambda^2$ practically the same as in pQCD) \cite{CCEM}. The remaining parameter, $M_0^2$, can be fixed by requiring that this coupling reproduce the experimental value of the massless strangeless semihadronic $\tau$-decay ratio $r_{\tau}=0.202\pm 0.004$, the low-energy observable that cannot be reproduced correctly using MA or mMA. {\noindent Work supported by Fondecyt grant 1095196 (G.C) and a PIIC-USM grant (H.M.).} \bibliographystyle{aipproc}
\section{Introduction} It is generally expected that wave packets evolving in a homogeneous random environment propagate diffusively over long time scales, unless recurrence effects are strong enough to induce Anderson localization. If furthermore the environment fluctuates in time, recurrence effects should be irrelevant, suggesting that diffusion is universal for wave motion in time dependent random systems. This idea was confirmed by Ovchinnikov and Erikman \cite {Ovchinnikov:1974eu}, who showed diffusion for a tight binding Schr\"odinger equation with white noise potentials. Pillet \cite{Pillet:1985oq} considered a more general setting in which the potentials were Markov processes, but not necessarily white noise. He demonstrated the absence of binding and derived a Feynman-Kac formula. This Feynman-Kac formula was used by Tcheremchantsev \cite{Tcheremchantsev:1997kl, Tcheremchantsev:1998qe} to show that position moments scale diffusively up to logarithmic corrections. Recently, two of us proved diffusion of wave packets and diffusive scaling \cite{KS} for a large family of Markov models, including those considered by Tcheremchantsev. For a recent discussion of the physics and physical applications of the tight binding Schr\"odinger equation with time dependent randomness we refer to \cite{Amir:2009}. The study of diffusion for disordered quantum systems, or ``Quantum Brownian motion," has recently attracted the attention of a number of authors. Diffusion in the presence of a weak static random potential for a quantum particle on a lattice of dimension three or higher has been demonstrated only up to a finite time scale proportional to an inverse power of the disorder strength \cite{Erdos2007:621, Erdos2007:1, Erdos2008:211}. Fr\"ohlich, Pizzo and De Roeck have proved diffusion, for arbitrarily long times, for a quantum particle on a lattice weakly coupled to an array of independent heat baths \cite{W.-De-Roeck:rz}. In \cite{W.-De-Roeck:rz}, it is mentioned that the method used therein also applies to a particle in a time dependent potential provided one has exponential decay of time correlations, such as one has for the Markov potentials in \cite{KS}. However the proof in \cite{W.-De-Roeck:rz} relies on a polymer expansion which restricts the result to weak coupling (this is not the case in \cite{KS}). Another result closely related to our previous work \cite{KS} is a recent paper on diffusion starting from a quantum master equation in Lindblad form \cite{Clark:2008}. This note and the aforementioned \cite{Pillet:1985oq,Tcheremchantsev:1997kl, Tcheremchantsev:1998qe,KS} are concerned with the evolution of wave packets for the ``tight binding Markov random Schr\"odinger equation:'' \begin{align}\label{Schrodinger} \begin{cases} \mathrm i \partial_{t} \psi_{t}(x) \ = \ L \psi_{t}(x) + v_{x}(\omega(t)) \psi_{t}(x), \\ \psi_{0} \in \ell^{2}(\mathbb Z^{d}), \end{cases} \end{align} where \begin{enumerate} \item $L$ is a translation invariant hopping operator on $\ell^{2}(\mathbb Z^{d})$, \item $v_{x}:\Omega \rightarrow \mathbb R$ are real valued functions on a probability space $\Omega$, \item $\omega(t)$ is a Markov process on $\Omega$ with an invariant probability measure $\mu$, and \item $v_{x}(\omega)=v_{0}(\sigma_{x}(\omega))$ where $\sigma_{x}$ is a group of $\mu$-measure preserving transformations of $\Omega$. \end{enumerate} (Formal definitions are given in section \ref{sec:main} below.) The potentials considered by Tcheremchantsev \cite{Tcheremchantsev:1997kl, Tcheremchantsev:1998qe} were independent at different sites. However, this played no role in the analysis in \cite{KS}. Nonetheless, some non-degeneracy assumption is certainly needed as can be seen by considering the case $v_{x}=v_{0}$ for all $x$, for which the effect of the random potential is only to multiply the wave function by a time dependent random phase. The technical condition employed in \cite{KS} was \begin{align} \inf_{x} \norm{B^{-1}(v_x - v_0)}_{L^{2}(\Omega)} >0 \label{v-non-degenerate}, \end{align} where $B$ is the generator of the Markov process $\omega(t)$. Our aim here is to consider a situation in which \eqref{v-non-degenerate} is violated in a relatively strong way. Namely, we shall consider \emph{periodic} potentials, $v_{x+Ny} = v_{x}$ for all $x,y$ with $N$ some fixed number. Because the resulting system is periodic under translations by elements of $N \mathbb Z^{d}$, there is a conserved ``quasi-momentum.'' Our main result, in short, is that after taking into account conservation of quasi-momentum the motion of the wave packet is diffusive. More specifically, over long times one sees a superposition of diffusions: \begin{equation}\label{eq:mainintro} \lim_{\tau \rightarrow \infty} \sum_{x \in \mathbb Z^{d}} e^{-\mathrm i \frac{1}{\sqrt{\tau}} \vec{k}\cdot x } \Ev{\abs{\psi_{\tau t}(x)}^{2}} \ = \ \int_{\mathbb{T}_{N}^{d}} e^{-t \sum_{i,j=1}^{d}D_{i,j}(\vec{p}) \vec {k}_{i}\vec {k}_{j}} m(\vec{p}) \mathrm d \vec{p}, \end{equation} where $\mathbb{T}_{N}^{d}=[0,2\pi/N)^{d}$, $\vec{p} \mapsto D_{i,j}(\vec{p})$ is a continuous function taking values in the positive definite matrices, independent of $\psi_{0}$, and \begin{equation}\label{eq:mintro} m(\vec{p}) \ = \ \frac{1}{(2 \pi )^{d}} \sum_{\zeta \in \Lambda} \abs{\wh{\psi}_{0}\left (\vec{p} + \frac{2\pi}{N} \zeta \right )}^{2} \end{equation} with $\Lambda =[0,N)^{d} \cap \mathbb Z^{d}$. The quantity $m(\vec{p})$ is the amplitude of the initial wave packet at quasi-momentum $\vec{p}$ \textemdash \ \ $\wh{\psi}_{0}$ denotes the Fourier transform of $\psi_{0}$: \begin{equation} \wh{\psi}_{0}(\vec{k}) \ =\ \sum_{x}\mathrm e^{\mathrm i x \cdot \vec{k}} \psi_{0}(x), \end{equation} if $\psi_{0} \in \ell^{1} \cap \ell^{2}$. To understand the meaning of \eqref{eq:mainintro}, consider the following position space density \begin{equation} \mathrm d R_{t}(x) = \sum_{\xi \in \mathbb Z^{d}} \Ev{|\psi_{t}(\xi)|^{2}} \delta(x-\xi)\mathrm d x, \end{equation} a probability measure on $\mathbb R^{d}$. (Here $\delta(x)\mathrm d x$ is the Dirac measure with mass $1$ at $0$.) After taking inverse Fourier transforms of both sides, \eqref{eq:mainintro} shows \begin{multline} \int_{\mathbb R^{d}} \phi(x) \mathrm d R_{\tau t}(\sqrt{\tau} x) \ \xrightarrow[\tau \rightarrow \infty]{} \\ \int_{\mathbb R^{d}} \phi(x) \left [\int_{\mathbb{T}_{N}^{d}} \frac{1}{(4 \pi t)^{\frac{d}{2}} \sqrt{\det D_{i,j}(\vec{p})}} \mathrm e^{-\frac{1}{4 t} \sum_{i,j} D_{i,j}^{-1}(\vec{p}) x_{i}x_{j}} m(\vec{p}) \mathrm d \vec{p} \right ]\mathrm d x, \end{multline} for any test function $\phi$ on $\mathbb R^{d}$ which is, say, smooth and compactly supported. The function appearing as the integrand inside square brackets on the right hand side is the fundamental solution to an anisotripic diffusion equation, with diffusion matrix $D_{i,j}(\vec{p})$, \begin{equation}\label{eq:diffusionequation} \frac{\partial}{\partial t} u_{t }(x) \ = \ \sum_{i,j} D_{i,j}(\vec{p})\frac{\partial}{\partial x_{i}} \frac{\partial}{\partial x_{j}} u_{t}(x). \end{equation} Thus \eqref{eq:mainintro} can be understood as saying that the position space density $dR_{t} (x)$, after diffusive rescaling $t \mapsto \tau t$ and $x \mapsto \sqrt{\tau}x$, converges in the weak$^{*}$ sense to \begin{equation} \mathrm d R_{\tau t}(\sqrt{\tau} x) \xrightarrow[\tau \rightarrow \infty]{\text{weak}^{*}} \left [ \int_{\mathbb{T}_{N}^{d}}u_{t}(x;\vec{p}) m(\vec{p}) \mathrm d \vec{p} \right ] \mathrm d x , \end{equation} where $u_{t}(x;\vec{p})$ satisfies \eqref{eq:diffusionequation} with $u_{0}(x;\vec{p}) \mathrm d x = \delta(x) \mathrm d x$. That is over long time scales, after diffusive rescaling, the mean square amplitude breaks into components for each $\vec{p}$, with each component propagating independently and according to a diffusion equation, which is to say a ``super-position of diffusions.'' The result is stated formally in section \ref{sec:main} after we give the required assumptions. These assumptions are somewhat technical, so it may be useful to have a simple example in mind. Fix a function $U:\mathbb Z^{d} \rightarrow \mathbb R$ periodic under translations in $N \mathbb Z^{d}$, that is, $U(x-Ny)=U(x)$ for all $x,y \in \mathbb Z^{d}$. Now let $\omega(t)$ be a continuous time random walk on $\Lambda =[0,N)^{d}\cap \mathbb Z^{d}$ taken with periodic boundary conditions and with independent identically distributed exponential holding times at each step. The probability space is just $\Lambda$ with the measure $\mu$ normalized counting measure. Take the potentials $v_{x}$ to be $v_{x}(\omega) = U(x-\omega)$ so that the Schr\"odinger equation describes a particle in a ``jiggling'' periodic potential: \begin{equation} \mathrm i \partial_{t} \psi_{t}(x) \ = \ \sum_{\zeta}h(\zeta) \psi_{t}(x-\zeta) + U(x-\omega(t)) \psi_{t}(x). \end{equation} Our result shows that \eqref{eq:mainintro} holds provided $U$ has no smaller periods, i.e. that $$\sum_{y \in \Lambda}|U(x+y)-U(y)| \neq 0 , \quad x \in \Lambda \text{ and }x \neq 0.$$ \section{Statement of the main result: A superposition of diffusions }\label{sec:main} \subsection{Assumptions} Our main result is formulated with the following assumptions. (See \cite{KS} for a more detailed discussion of the framework.) \begin{asst} We are given a topological space $\Omega$, a Borel probability measure $\mu$, and a Markov process on $\Omega$ with right continuous paths for which $\mu$ is an invariant measure. Furthermore, we suppose that there is a representation of $\mathbb Z^{d}$, $x \mapsto \sigma_{x}$, in terms of $\mu$-measure preserving maps $\sigma_{x}:\Omega \rightarrow \Omega$ such that the paths of $\sigma_{x}(\omega(\cdot))$ have the same distribution as the paths of $\omega(\cdot)$, for all $x \in \mathbb Z^{d}$. \end{asst} We denote by $\Ev{\cdot}$ expectation with respect to the paths of the Markov process with the initial condition $\omega(0)$ distributed according to $\mu$. By the invariance of $\mu$, we have \begin{equation}\label{eq:invariantexp} \Ev{f(\omega(t))} = \int_{\Omega}f(\alpha) \mathrm d \mu(\alpha) \end{equation} for any $t$ and any $f \in L^{1}(\Omega)$. Furthermore, the map $S_{t}$ given by \begin{equation} S_{t}f(\alpha) \ = \ \mathbb E( f(\omega(0)) | \omega(t) =\alpha) \end{equation} defines a strongly continuous contraction semi-group on $L^{2}(\Omega)$. By the Lumer-Phillips theorem, $S_{t}$ is generated by a maximally dissipative operator $B$ with dense domain $\cu{D}(B)$. Since $S_{t}1 =1$ for all $t$, $B1=0$ and $0$ is an eigenvalue of $B$. Since $B$ is dissipative, we also have that its numerical range lies in the right half plane. We suppose further that $B$ is sectorial and satisfies a ``spectral gap'' condition: \begin{asst} There exist $\gamma < \infty$ and $T >0$ such that \begin{align} \abs{\operatorname{Im} \ip{f, B f}_{L^{2}(\Omega)}} \ & \le \ \gamma \operatorname{Re} \ip{f, B f}_{L^{2}(\Omega)} , \quad \label{eq:sectoriality} \intertext{and} \operatorname{Re}\ip{f, B f}_{L^{2}(\Omega)} & \ge \frac{1}{T} \operatorname{Var} ( f) \label{eq:spectralgap} \end{align} for all $f \in \cu{D}(B)$, where $\operatorname{Var}(f):=\int_{\Omega} f^{2}\mathrm d \mu - (\int_{\Omega} f \mathrm d \mu)^{2}$.\end{asst} The potential $v_{x}:\Omega \rightarrow \mathbb R$ and hopping operator $L$ are assumed to be translation invariant, and $L$ should satisfy a non-degeneracy condition that precludes hopping only in a sub-lattice: \begin{asst} The potential is given by Borel measurable bounded functions $v_{x} :\Omega \rightarrow \mathbb R$ such that $$ \quad v_{x} = v_{0} \circ \sigma_{x} . $$ The hopping operator is given by $$L \psi(x) \ = \ \sum_{y} h(x-y) \psi(y),$$ where $h(-x)=h(x)^{*}$, $ \sum_{x} |x|^{2} \abs{h(x)} < \infty$, and for each non-zero vector $\vec k \in \mathbb R^{d}$, there is some $x \in \mathbb Z^{d}$ such that $h(x) \neq 0$ and $\vec k \cdot x \ne 0$. \end{asst} Finally, since we are concerned with periodic potentials, we suppose \begin{asst} There is $N \in \mathbb N$, $N > 1$, such that $\sigma_{Nx}= \operatorname{Id}$ for all $x \in \mathbb Z^{d}$. Furthermore, we suppose that $\norm{v_{x}-v_{0}}_{L^{\infty}(\Omega)}>0$ for all $x \in [0,N)^{d}\cap \mathbb Z^{d}$, $x \neq 0$. \end{asst} \begin{rem*} More generally, we might allow different periods in each of the coordinate directions: $N_{1}, \ldots N_{d}$ such that $\sigma_{y} = \operatorname{Id}$ whenever $y = (N_{1} \alpha_{1}, \ldots, N_{d}\alpha_{d})$ with $\alpha_{1}, \ldots, \alpha_{d} \in \mathbb Z$. The result stated below holds also for this case with essentially the same proof. We choose to work with equal periods for notational clarity. \end{rem*} Let $\Lambda = [0,N)^d \cap \mathbb Z^{d}$, as above, and let $x,y\in\Lambda$. Since $v_{x}-v_{y}$ is mean zero, it is in the domain of $B^{-1}$. Furthermore, it follows from Assumption 4 that $\norm{v_{x}-v_{y} }_{L^{2}(\Omega)}\neq 0$ if $x\neq y$, in which case $\norm{B^{-1}(v_{x}-v_{y})}_{L^{2}(\Omega)}\neq 0$. Since $\Lambda$ is finite, we conclude that there is $\chi > 0$ such that \begin{equation}\label{eq:nondeg} \norm{B^{-1} (v_{x}- v_{y})}_{L^{2}(\Omega)} \ge \chi \ , \quad x,y \in \Lambda , \ x \neq y. \end{equation} Eq. \eqref{eq:nondeg} will play a key role in the proof below. \subsection{Main result} Consider the density matrix \begin{equation} \rho_{t}(x,y) \ = \ \psi_{t}(x) \psi_{t}(y)^{*}. \end{equation} It is well-known that $\rho_t(x,y)$ satisfies \begin{equation}\label{MAMDM} \partial_{t} \rho_{t}(x,y) \ = \ - \mathrm i \sum_{\zeta} h(\zeta) \left[ \rho_{t}(x-\zeta,y) - \rho_{t}(x,y+\zeta) \right ] -\mathrm i \left ( v_{x}(\omega(t)) - v_{y}(\omega(t)) \right ) \rho_{t}(x,y). \end{equation} More generally, we may consider solutions to \eqref{MAMDM} with an initial condition \begin{multline} \rho_{0} \in \cu{DM} \ := \ \left \{ \rho: \mathbb Z^{d} \times \mathbb Z^{d} \rightarrow \mathbb C \ : \ \rho \text{ is the kernel of a non-negative definite,} \right . \\ \left. \text{trace class operator on $\ell^{2}(\mathbb Z^{d})$} \right \}. \end{multline} Recalling the notation $\mathbb{T}_{N}^{d}=[0,2\pi/N)^{d}$, we now state our theorem. \begin{thm}\label{thm:main} The solution to \eqref{MAMDM} with initial condition $\rho_{0} \in \cu{DM}$ satisfies \begin{equation}\label{eq:main} \lim_{\tau \rightarrow \infty }\sum_{x} \mathrm e^{-\mathrm i \frac{\vec{k}}{\sqrt{\tau}} \cdot x} \Ev{\rho_{\tau t} (x,x)} \ = \ \int_{\bb{T}_{N}^d} \, \mathrm e^{-t \sum_{i,j}D_{i,j}(\vec p)\vec{k}_{i} \vec{k}_{j}} m(\vec p)\mathrm d \vec{p}, \end{equation} where $\vec{p} \mapsto D_{i,j}(\vec p)$ is a continuous function taking values in the positive-definite matrices and $$m(\vec{p}) \ = \ \frac{N^{d}}{(2\pi)^{d}} \wh{f}(N\vec{p}) , $$ with $\wh{f}$ the Fourier transform of $f(x) = \sum_{y \in \mathbb Z^{d}} \rho_{0}(y+Nx,y) .$ \end{thm} \begin{rem*}We have defined the function $m$ in terms of the Fourier transform of $f$. Since $f$ is not obviously summable or square summable, it is not immediately clear that $m$ is indeed a \emph{function}, rather than a distribution. However, in terms of the orthnormal eigenvectors $\psi_{j}$ of $\rho_{0}$ and corresponding eigenvalues $\lambda_{j}$, we have \begin{equation}\label{eq:nobochner} m(\vec{p}) \ = \ \frac{1}{(2\pi)^{d}} \sum_{j} \lambda_{j} \sum_{\zeta \in \Lambda} \abs{\wh{\psi_{j}}\left (\vec{p} + \frac{2 \pi}{N} \zeta \right )}^{2}. \end{equation} Since $\sum_{j} \lambda_{j} <\infty$ and $\abs{\wh{\psi_{j}}}^{2} \in L^{1}(\mathbb{T}^{d})$ we see that $m(\vec{p})$ is an $L^{1}$ function of $\vec{p}$. (The function $f(x)$ can be expressed as \begin{equation} f(x) \ = \ \tr \rho_{0} S_{Nx} \end{equation} where $\rho_{0}$ is interpreted as a trace class operator and $S_{Nx}$ is the shift by $Nx$ on $\ell^{2}(\mathbb Z^{d})$, $S_{Nx}\psi(y)=\psi(y-Nx)$. It follows that $f$ is \emph{positive definite}: \begin{equation} \sum_{i,j=1}^{n}\zeta_{i}^{*}\zeta_{j}f(x_{i}-x_{j})\ge 0 \end{equation} for any finite collection of points $x_{1},\ldots, x_{n} \in \mathbb Z^{d}$ and any $(\zeta_{1},\ldots, \zeta_{n})\in \mathbb C^{n}$. We conclude from Bochner's theorem that $\wh{f}$ is a non-negative measure of mass $f(0)=\tr\rho_{0}$, and because $\lim_{x\rightarrow \infty}f(x)=0$ the measure has no point component. But, it is not immediately clear that $\wh{f}$ is absolutely continuous with respect to Lebesuge measure so that $m$ is a function. For this purpose \eqref{eq:nobochner} seems to be necessary.) \end{rem*} \section{Augmented space analysis} In this section, we explain briefly the augmented space analysis, which was also employed in \cite[Section 3]{KS}. We begin with the following Feynman-Kac formula \cite{Pillet:1985oq} \begin{equation} \mathbb E(\rho_t (x,y)) = \ip{ \delta_{x}\otimes \delta_{y} \otimes 1, e^{-t L } \rho_0 \otimes 1}_{\cu{H}}, \label {E(Rho)} \end{equation} which relates $\Ev{\rho_{t}(x,x)}$ to a matrix element of a contraction semigroup $e^{-tL}$ on the augmented Hilbert space \begin{equation} \cu{H} \ := \ L^{2}(\mathbb Z^{d} \times \mathbb Z^{d} \times \Omega). \end{equation} The operator $L$ in \eqref{E(Rho)} is given by $ L := \mathrm i K + \mathrm i V +B$, where \begin{align} K \Psi(x,y,\omega) \ &= \ \sum_{\zeta} h(\zeta) \left[ \Psi(x-\zeta,y,\omega) - \Psi(x,y+\zeta,\omega) \right ], \\ V \Psi(x,y,\omega) \ &= \ \left ( v_{x}(\omega) -v_{y}(\omega) \right ) \Psi(x,y,\omega). \end{align} The Markov generator $B$ acts on $\cu{H}$ as a multiplication operator with respect to the first two coordinates: \begin{equation} B [\rho \otimes f] \ = \ \rho \otimes (Bf), \quad \rho \in \ell^{2}(\mathbb Z^{d} \times \mathbb Z^{d}) , \ f \in L^{2}(\Omega). \end{equation} Our analysis, as in \cite{KS}, makes crucial use of the invariance of the generator $L$ with respect to simultaneous translation of position and disorder. In the present context, we have a larger group of symmetries due to periodicity. Namely, the generator $L$ and its constituents $K$, $V$, and $B$, commute with a group $\cu{G}$ of unitary maps on $\cu{H}$ generated by the following transformations: \begin{enumerate} \item Simultaneous translation of position and disorder by an arbitrary element of $\mathbb Z^{d}$: $$ S_{\xi}\Psi(x,y,\omega)=\Psi(x-\xi,y-\xi,\sigma_{\xi}\omega),$$ \item Translation of the first position coordinate by an element of $N \mathbb Z^{d}$: $$ S^{(1)}_{N \xi}\Psi(x,y,\omega)=\Psi(x-N\xi, y, \omega).$$ \end{enumerate} Note that $S_{\xi}S^{(1)}_{N\eta}=S^{(1)}_{N\eta}S_{\xi}$, so the group $\cu{G}$ is isomorphic to $\mathbb Z^{d}\times \mathbb Z^{d}$. We have chosen to use translation of the first position in the definition of $S^{(1)}$; however, since $\sigma_{N\xi}=\operatorname{Id}$, we have $S^{(2)}_{N\xi}=S_{N\xi}S^{(1)}_{-N\xi} \in \cu{G}$, where $ S^{(2)}_{N \xi}\Psi(x,y,\omega)=\Psi(x,y-N\xi, \omega).$ Because of the invariance with respect to $\cu{G}$, $L$ is partially diagonalized by the following generalized Fourier transform: \begin{equation} \widetilde{ \Psi}(x,\omega, \vec{k}, \vec{p}) = \sum_{\xi,\eta \in \mathbb Z^{d} } \mathrm e^{\mathrm i \vec{p} \cdot (x-N\eta) -\mathrm i \vec{k}\cdot \xi} \Psi (x -\xi -N\eta , -\xi , \sigma_{\xi}\omega), \end{equation} a unitary map from $L^2(\mathbb Z^{d} \times \mathbb Z^{d} \times \Omega) \to L^2(\Lambda \times \Omega \times \bb{T}_{1}^{d} \times \bb{T}_{N}^d).$ Thus we have, by \eqref{E(Rho)}, \begin{equation} \sum_{x} \mathrm e^{-\mathrm i \vec{k} \cdot x} \Ev{\rho_{t}(x,x)} \ = \ \frac{N^{d}}{(2\pi)^{d}}\int_{ \bb T_{N}^d} \mathrm d \vec{p} \ip{ \delta_{0} \otimes 1, \mathrm e^{-t \widetilde{L}_{\vec{k} , \vec{p}}} \widetilde{\rho}_{0;\vec{k}, \vec{p}}\otimes 1}_{L^{2}(\Lambda \times \Omega)} \label {E(Rho3)FT}, \end{equation} where \begin{align} \widetilde{\rho}_{0;\vec{k},\vec{p}}(x) \ &= \ \sum_{y, \eta} \mathrm e^{\mathrm i \vec{p}\cdot(x-N \eta)-\mathrm i \vec{k}\cdot y} \rho_{0}(x-N\eta -y,-y), \end{align} and $\widetilde{L}_{\vec{k} , \vec{p}} \ := \ \mathrm i \widetilde{K}_{\vec{k} , \vec{p}} + \mathrm i \widetilde{V} + B$ with \begin{align} \widetilde{V} \widetilde{\psi}(x,\omega) &= (v_{x}(\omega) -v_{0}(\omega))\widetilde \psi (x, \omega), \intertext{and} \widetilde{K}_{\vec{k}, \vec p} \widetilde{\psi}(x,\omega ) &= \sum_{\zeta} h(\zeta)\mathrm e^{\mathrm i \vec {p} \cdot \zeta} \left [ \widetilde{\psi}(x-\zeta,\omega) - \mathrm e^{-\mathrm i \vec{k} \cdot \zeta} \widetilde{\psi}(x-\zeta, \sigma_{\zeta}\omega) \right]. \label{eq:wtK} \end{align} (In \eqref{eq:wtK} we take ``periodic boundary conditions,'' that is $x -\zeta$ on the right hand side is evaluated modulo $N$.) The transformed Feynmann-Kac formula \eqref{E(Rho3)FT} is the starting point for our proof of Theorem \ref{thm:main}. It reduces the study of the mean density in \eqref{eq:main} to the spectral analysis of the semi-group $\mathrm e^{-t \widetilde{L}_{\vec{k}, \vec{p}}}$ for each fixed $\vec p$ and for $\vec{k}$ in a small neighborhood of $0$. \section{Spectral analysis of $\widetilde{L}_{\vec{k} , \vec{p}}$ and the proof of Theorem \ref{thm:main}} In this section inner products and norms are taken in the space $L^{2}(\Lambda \times \Omega)$ unless otherwise indicated. We denote by $P_{0}$ the orthogonal projection of $L^{2}(\Lambda \times \Omega)$ onto the space $\cu{H}_{0}= \ell^{2}(\Lambda)\otimes \{1\}$ of ``non-random'' functions, \begin{equation} P_{0}\Psi(x) \ = \ \int_{\Omega } \Psi(x,\omega) \mathrm d \mu(\omega), \end{equation} and by $P_{0}^{\perp}=(1-P_{0})$ the projection onto mean zero functions \begin{equation} \cu{H}^{\perp}_{0} = \set{ \Psi(x,\omega) \ : \ \int_{\Omega} \Psi(x,\omega) \mathrm d\mu(\omega) = 0}. \end{equation} A preliminary observation is that \begin{equation}\label{eq:groundstate} \widetilde{L}_{\vec{0},\vec{p}} \delta_{0} \otimes 1 = 0 \end{equation} for all $\vec{p}$. Thus, $\delta_{0} \otimes 1$ is stationary under each semigroup $e^{-t \widetilde{L}_{\vec{0},\vec{p}}}$. Eq.\ \eqref{eq:groundstate} can be seen easily from the explicit form for $\widetilde{L}_{\vec{0},\vec{p}}$ given above, but could also be derived from the fact that, for each $y \in \mathbb Z^{d}$, $$\sum_{x} \Ev{\rho_{t}(x+Ny,x)}$$ is constant in time. A key step toward proving Theorem \ref{thm:main} is to observe that the remaining spectrum of $\widetilde{L}_{\vec{0},\vec{p}}$ is contained in a half plane with strictly positive real part. To see this, we make use of the block decomposition of $\widetilde{L}_{ \vec{0}, \vec{p}}$ with respect to the direct sum $\cu{H}_{0} \oplus \cu{H}_{0}^{\perp}$: \begin{equation}\label{eq:blockform} \widetilde{L}_{\vec{0} , \vec{p}} \ = \ \begin{pmatrix} 0 & \mathrm i P_{0} \widetilde{V} \\ \mathrm i \widetilde{V} P_{0} & \mathrm i \widetilde{K}_{\vec{0}, \vec{p}} + B + \mathrm i P_{0}^{\perp} \widetilde{V} P_{0}^{\perp} \end{pmatrix}. \end{equation} (Note that $\widetilde{K}_{\vec{0},\vec{p}}$ and $B$ both act trivially on $\cu{H}_{0}$, while $P_{0} \widetilde{V} P_{0}=0$ since $\int_{\Omega}(v_{x}(\omega) -v_{0}(\omega) ) d \mu(\omega)=0$.) We use \eqref{eq:blockform} to prove the following \begin{lem}\label{lem:L0} There is $\delta >0$ such that for all $\vec{p} \in \bb{T}^{d}_{N}$, \begin{equation} \sigma(\widetilde{L}_{\vec{0}, \vec{p}}) \ = \ \{0\} \cup \Sigma_{+}(\vec{p}) \end{equation} where $0$ is a non-degenerate eigenvalue and $\Sigma_{+}(\vec{p}) \subset \set{z \ : \ \operatorname{Re} z >\delta}.$ \end{lem} \begin{proof} This is very close to \cite[Lemma 3]{KS}. The key new point is that we must see that $\delta$ can be chosen independently of $\vec{p}$. Because $\operatorname{Re} B \ge \frac{1}{T} P_{0}^{\perp}$, it follows from an argument using Schur complements that a point $z$ with $\operatorname{Re} z < \frac{1}{T}$ is in $\sigma(\widetilde{L}_{\vec{0},\vec{p}})$ if and only if $z$ is in the spectrum of \begin{equation} \Gamma_{\vec{p}}(z) = P_{0} \widetilde{V} (P_{0}^{\perp} \widetilde{L}_{\vec{0},\vec{p}} P_{0}^{\perp}-z)^{-1} \widetilde{V} P_{0}. \end{equation} However, given $\phi \in \ell^{2}(\Lambda)$, \begin{align} \operatorname{Re} & \ip{\phi\otimes 1, \Gamma_{\vec{p}}(z) \phi \otimes 1} \notag \\ &= \left \langle (P_{0}^{\perp} \widetilde{L}_{\vec{0} , \vec{p}} P_{0}^{\perp}-z)^{-1} \widetilde{V} \phi \otimes 1 \, ,\, (\operatorname{Re} B - \operatorname{Re} z) (P_{0}^{\perp} \widetilde{L}_{\vec{0},\vec{p}} P_{0}^{\perp}-z)^{-1} \widetilde{V} \phi \otimes 1 \right \rangle \notag \\ & \ge \left ( \frac{1}{T} - \operatorname{Re} z \right ) \norm{ (B^{-1} P_{0}^{\perp} (\widetilde{L}_{\vec{0},\vec{p}} -z )P_{0}^{\perp})^{-1} B^{-1} \widetilde{V} \phi \otimes 1}^{2}, \end{align} where the inverses are well defined because $\widetilde{V} \phi \otimes 1 \in \cu{H}_{0}^{\perp}=\operatorname{ran} P_{0}^{\perp}$. Since $\norm{B^{-1} P_{0}^{\perp}} \le T$, it follows that \begin{equation} \norm{B^{-1} P_{0}^{\perp} (\widetilde{L}_{\vec{0},\vec{p}} -z )P_{0}^{\perp} } \ \le \ 1 + T\left (\| \widetilde{K}_{\vec{0},\vec{p}} \| + \| \widetilde{V} \|+ |z| \right ). \end{equation} However, $\|\widetilde{K}_{\vec{k},\vec{p}}\|\le 2 \|\wh{h}\|_{\infty}$ for all $\vec{k}$ and $\vec{p}$, so $B^{-1} P_{0}^{\perp} (\widetilde{L}_{\vec{0},\vec{p}} -z )P_{0}^{\perp}$ is uniformly bounded and \begin{align} \operatorname{Re} & \ip{\phi\otimes 1, \Gamma_{\vec{p}}(z) \phi \otimes 1} \notag \\ & \ge \left ( \frac{1}{T} - \operatorname{Re} z \right ) \frac{1}{ \left [1 + T(2 \| \wh{h}\|_{\infty} + 2\|\widetilde{V}\| + |z|) \right ]^{2}} \norm{ B^{-1} \widetilde{V} \phi \otimes 1}^{2}. \end{align} Finally, \begin{equation} \norm{ B^{-1} \widetilde{V} \phi \otimes 1}^{2} \ = \ \sum_{x}|\phi(x)|^{2} \norm{B^{-1}(v_{x}-v_{0})}_{L^{2}(\Omega)}^{2} \ \ge \ \chi^{2} \sum_{x \neq 0}|\phi(x)|^{2}, \end{equation} where \begin{equation} \chi \ = \ \min_{\substack{x \in \Lambda \\ x \neq 0}} \norm{B^{-1}(v_{x}-v_{0})}_{L^{2}(\Omega)}, \end{equation} which is positive by Assumption 4. Thus, \begin{equation} \operatorname{Re} \Gamma_{\vec{p}}(z) \ \ge \ \left ( \frac{1}{T} - \operatorname{Re} z \right ) \frac{\chi^{2}}{\left [1 + T(2 \| \wh{h}\|_{\infty} + 2\|\widetilde{V}\| + |z|) \right ]^{2}}. \end{equation} Since the right hand side is independent of $\vec{p}$, the existence of a spectral gap $\delta$ independent of $\vec{p}$, as claimed, now follows from the sectoriality of $B$ (Assumption 2, eq.\ \eqref{eq:sectoriality}) as in the proof of \cite[Lemma 3]{KS}, with the explicit estimate \begin{equation} \delta \ge \frac{1}{T} \frac{\chi^{2 }}{\left ( 2 + \gamma + 4 T \|\wh{h}\|_{\infty} + 4 T \|\widetilde{V}\|\right )^{2}+ \| \widetilde{V}\|^{2}\chi^{2}}. \qedhere \end{equation} \end{proof} \subsection{Analytic perturbation theory for $\widetilde{L}_{\vec{k} , \vec{p}}$} We now hold $\vec{p}$ fixed and consider the spectrum of $\widetilde{L}_{\vec{k},\vec{p}}$ for $\vec{k}$ close to $0$. We write $\nabla$ for the gradient with respect to $\vec{k}$ and $\partial_{i}$ for partial differentiation with respect to the $i^{\text{th}}$ coordinate of $\vec{k}$. \emph{No derivatives with respect to $\vec{p}$ appear below}. The key observation is that the spectral gap for $\widetilde{L}_{\vec{0}, \vec{p}}$ is preserved in the spectrum of $\widetilde{L}_{\vec{k} , \vec{p}}$ for $\vec{k}$ sufficiently small. \begin{lem}\label{lem:analytic} Given $\epsilon \in (0,\delta)$, with $\delta$ as in Lemma \ref{lem:L0}, there exists $r$ such that if $|\vec{k}| < r $ then, for each $\vec{p} \in \bb{T}^{d}_{N}$, \begin{enumerate} \item $\widetilde{L}_{\vec{k} , \vec{p}}$ has a single non-degenerate eigenvalue $E_{\vec {p}}(\vec k)$ with $0 \le \operatorname{Re} E_{\vec{p}}(\vec{k}) < \delta -\epsilon$, \item The rest of the spectrum of $ \widetilde{L}_{\vec{k} , \vec{p}}$ is contained in the half plane $ \{ z : \operatorname{Re} z > \delta - \epsilon \}$. \end{enumerate} Furthermore, $E_{\vec {p}}(\vec{k})$ is $C^{2}$ in a neighborhood of $0$, \begin{equation}\label{eq:explicit0} E_{\vec {p}}(\vec 0) = 0, \quad \nabla E_{\vec {p}}(\vec 0) = 0, \end{equation} and \begin{multline}\label{eq:explicit} \partial_{i} \partial_{j} E_{\vec {p}}(\vec 0) \ = \ 2 \operatorname{Re} \ip{ \partial_{i} \widetilde{K}_{\vec{0} , \vec{p}} \delta_{0} \otimes 1, [\widetilde{L}_{\vec{0} , \vec{p}}]^{-1} \partial_{j} \widetilde{K}_{\vec{0} , \vec{p}} \delta_{0} \otimes 1} \\ = \ 2 \operatorname{Re} \sum_{x,y \in \mathbb Z^{d}} x_{i} y_{j} \overline{h(x)} h(y) \ip{ \delta_{[x]_{N}} \otimes 1 ,[\Gamma_{\vec{p}}(0)]^{-1} \delta_{[y]_{N}}\otimes 1} , \end{multline} where $[x]_{N}$ denotes the point in $\Lambda$ equivalent to $x$ modulo $N$ and \begin{equation} \Gamma_{\vec{p}}(0) = P_{0}\widetilde{V} ( P_{0}^{\perp} \widetilde{L}_{\vec{0} , \vec{p}}P_{0}^{\perp} )^{-1} \widetilde{V} P_{0}. \end{equation} In particular, $\partial_{i} \partial_{j} E_{\vec {p}}(\vec 0)$ is positive definite. \end{lem} \begin{proof} These are essentially standard facts from analytic perturbation theory. The key point is that \begin{equation} \norm{\widetilde{L}_{\vec{k} , \vec{p}} - \widetilde{L}_{\vec{0} , \vec{p}}} \ \le \ c |\vec{k}|. \label{difference_in_k} \end{equation} If the generators $\widetilde{L}_{\vec{k},\vec{p}}$ were self-adjoint or normal it would now follow that the spectrum moves by no more than a distance $c |\vec{k}|$ for $\vec{k}$ small. However, $\widetilde{L}_{\vec{k},\vec{p}}$ need not be normal so we must argue more carefully. Due to the spectral gap $\delta$ between $0$ and the rest of the spectrum of $\widetilde{L}_{\vec{0},\vec{p}}$, we can fit a contour $\cu{C}$ around the origin in the resolvent set. Then \eqref{difference_in_k} shows that the spectrum cannot cross $\cu{C}$ for small $\vec{k}$. A convenient choice for $\cu{C}$ is the rectangle $$ \cu{C}= \left ( \delta -\epsilon + \mathrm i [-R,R] \right ) \cup \left ( [-R, \delta -\epsilon] + \mathrm i R \right ) \cup \left ( -R + \mathrm i [-R,R] \right ) \cup\left ( [-R, \delta -\epsilon] - \mathrm i R \right ),$$ with $R$ fixed independent of $\epsilon$, but sufficiently large. By Lemma \ref{lem:L0}, \begin{equation} \label{boundonC} \sup_{\substack{z \in \cu{C} \\ \vec{p} \in \bb{T}^{d}_{N}}}\norm{ (\widetilde{L}_{\vec{0},\vec{p}} -z)^{-1}} < \infty. \end{equation} Expanding the resolvent of $\widetilde{L}_{\vec{k},\vec{p}}$ in a Neumann series, \begin{equation} ( \widetilde{L}_{\vec{k},\vec{p}} - z )^{-1} = \sum_{n=0}^{\infty } ( \widetilde{L}_{\vec{0},\vec{p}} - z )^{-1} \left [ (\widetilde{L}_{\vec{0} , \vec{p}} - \widetilde{L}_{\vec{k} , \vec{p}}) ( \widetilde{L}_{\vec{0},\vec{p}} - z )^{-1}\right ]^{n}, \end{equation} and using \eqref{difference_in_k} and \eqref{boundonC}, we see that there is $r > 0$ such that if $|\vec{k}| < r$, then $\cu{C}$ is in the resolvent set of $\widetilde{L}_{\vec{k},\vec{p}}$. However, the spectrum is a subset of the numerical range and the numerical range of $\widetilde{L}_{\vec{k},\vec{p}}$ is contained in the set \begin{equation} \set{ x+ \mathrm i y \ : \ x \ge 0 \ \& \ | y| \le C + \gamma x}, \end{equation} with $C = 2 \| \wh{h}\|_{\infty} + 2 \|\widetilde{V}\|$. We conclude that \begin{equation} \sigma(\widetilde{L}_{\vec{k},\vec{p}}) = \Sigma_{0} \cup \Sigma_{1} \end{equation} with $\Sigma_{0}$ inside $\cu{C}$ and $\Sigma_{1} \subset \{ z \ : \ \operatorname{Re} z > \delta -\epsilon\}$. It remains to show that $\Sigma_{0}$ consists of a non-degenerate eigenvalue and to derive \eqref{eq:explicit0} and \eqref{eq:explicit}. For this purpose, consider the (non-Hermitian) Riesz projection \begin{equation}\label{eq:Qk} Q_{\vec{k},\vec{p}} \ = \ \frac{1}{2 \pi \mathrm i} \int_{\cu{C}} \frac{1}{z - \widetilde{L}_{\vec{k} , \vec{p}}} \mathrm d z. \end{equation} The rank of $Q_{\vec{k},\vec{p}}$ is constant so long as $\cu{C}$ remains in the resolvent set. Thus, $Q_{\vec{k} , \vec{p}}$ is rank one for $|\vec{k}| <r$ and $\Sigma_{0} = \{ E_{\vec{p}}(\vec{k})\}$ with associated normalized eigenvector $\Phi_{\vec{k},{\vec {p}}}$ in the one-dimensional range of $Q_{\vec{k},\vec{p}}$. Then, $E_{\vec {p}}(\vec 0)=0$ and $\Phi_{\vec 0 , \vec{p}}= \delta_{0}\otimes 1$. By the Feynman-Hellman formula, \begin{equation} \partial_{i} E_{\vec {p}}(\vec{k}) \ = \ \ip{\Phi_{\vec{k},\vec{p} }, \partial_{i} \widetilde{L}_{\vec{k} , \vec{p}} \Phi_{\vec{k},\vec{p}}}, \end{equation} from which it follows that $\nabla E_{\vec {p}}(\vec 0) =0$ since $\nabla \widetilde{L}_{\vec{k} , \vec{p}}= \mathrm i \nabla \widetilde{K}_{\vec{k} , \vec{p}}$ is off-diagonal in the position basis on $\cu{H}_{0}$. Similarly, \begin{multline} \partial_{i} \partial_{j} E_{\vec {p}}(\vec{k}) \ = \ \ip{ \Phi_{\vec{k}, \vec{p}}, \partial_{i}\partial_{j} \widetilde{L}_{\vec{k} , \vec{p}} \Phi_{\vec{k}, \vec{p}}} + \ip{\Phi_{\vec{k}, \vec{p}}, Q_{\vec{k}} \partial_{i} \widetilde{L}_{\vec{k} , \vec{p}} ( E_{\vec{p}}(\vec{k}) - \widetilde{L}_{\vec{k} , \vec{p}})^{-1} (1-Q_{\vec{k}, \vec{p}}) \partial_{j} \widetilde{L}_{\vec{k} , \vec{p}} \Phi_{\vec{k}, \vec{p}}} \\ + \ip{\Phi_{\vec{k}, \vec{p}}, Q_{\vec{k}} \partial_{j} \widetilde{L}_{\vec{k} , \vec{p}} ( E_{\vec{p}}(\vec{k}) - \widetilde{L}_{\vec{k} , \vec{p}})^{-1} (1-Q_{\vec{k}, \vec{p}}) \partial_{i} \widetilde{L}_{\vec{k} , \vec{p}} \Phi_{\vec{k}, \vec{p}}} \end{multline} The first term on the r.h.s.\ vanishes at $\vec{k}=0$ and the remaining two terms give \eqref{eq:explicit}. Because the form on the r.h.s of \eqref{eq:explicit} is positive definite, the non-degeneracy condition on the hopping (Assumption 3) gives that $\partial_{i} \partial_{j} E_{\vec {p}}(\vec{0})$ is positive definite. \end{proof} It follows from Lemma \ref{lem:analytic} and the sectoriality \eqref{eq:sectoriality} of $B$ that the semigroup $e^{-t \widetilde{L}_{\vec{k},\vec{p}}}$ satisfies exponential bounds (see \cite[Lemma 4]{KS}): \begin{lem}\label{lem:Lkdynamics} Given $\epsilon >0$ there is $C_{\epsilon } <\infty$ such that if $\vec{k}$ is sufficiently small, then \begin{equation} \norm{\mathrm e^{-t \widetilde{L}_{\vec{k} , \vec{p}}}(1- Q_{\vec{k} , \vec{p}})} \le C_{\epsilon}\mathrm e^{-t (\delta - \epsilon )} \end{equation} for all $\vec{p}$, where $Q_{\vec{k},\vec{p}}$ is the rank one Riesz projection \eqref{eq:Qk} onto the non-degenerate eigenvector of $\widetilde{L}_{\vec{k},\vec{p}}$ with eigenvalue near $0$. \end{lem} \subsection{Proof of Theorem \ref{thm:main}} As in \cite{KS}, it suffices to prove the theorem for $\rho_{0}$ satisfying \begin{equation}\label{eq:summable} \sum_{x,y} |\rho_{0}(x,y)| < \infty, \end{equation} since any initial density matrix can be approximated in trace norm arbitrarily well using such $\rho_{0}$. Assuming \eqref{eq:summable}, note that \begin{equation} \widetilde{\rho}_{0;\vec{k}, \vec{p}}(x) \ = \ \sum_{\eta,y \in \mathbb Z^d} \rho_{0}(x-N\eta-y ,-y)\mathrm e^{\mathrm i \vec{p}\cdot (x-N \eta )-\mathrm i \vec{k}\cdot y} \label{fourier_transform_rho} \end{equation} is uniformly bounded in $\ell^{2}(\Lambda)$ as $\vec{p}$ varies through the torus: \begin{equation} \left [ \sum_{x}\abs{\widetilde \rho_{0;\vec{k}, \vec{p}}(x)}^{2} \right ]^{\frac{1}{2}} \ \le \ \sum_{x}\abs{\widetilde \rho_{0;\vec{k}, \vec{p}}(x)} \ \le \ \sum_{x,y} \abs{\rho_{0}(x,y)} < \infty. \end{equation} By \eqref{E(Rho3)FT}, we have \begin{align} \label{eq:projectionsdecompose} \sum_{x} \mathrm e^{-\mathrm i \frac{1}{\sqrt{\tau}} \vec{k} \cdot x} \Ev{\rho_{\tau t}(x,x)} \ & = \ \frac{N^{d}}{(2\pi)^{d}}\int_{ \bb{T}_{ N}^d}\mathrm d \vec{p}\ip{ \delta_{0} \otimes 1, \mathrm e^{-\tau t \widetilde{L}_{ \vec{k}/\sqrt{\tau}, \vec{p}}} \widetilde{\rho}_{0;\frac{1}{\sqrt{\tau}}\vec{k}, \vec {p}}\otimes 1} \\ & = \ \frac{N^{d}}{(2\pi)^{d}} \int_{\bb{T}_{ N}^d}\mathrm d\vec{p} \ \mathrm e^{-\tau t E_{\vec {p}}(\vec{k}/\sqrt{\tau})} \ip{\delta_{0}\otimes 1, Q_{\frac{1}{\sqrt{\tau}}\vec{k},\vec{p}} \widetilde{\rho}_{0;\frac{1}{\sqrt{\tau}} \vec{k}, \vec {p}} \otimes 1} \label {1st_term}\\ & \quad + \ \frac{N^{d}}{(2\pi)^{d}} \int_{\bb{T}_{ N}^d}\mathrm d\vec{p} \ \ip{ \delta_{0} \otimes 1, \mathrm e^{-\tau t \widetilde{L}_{ \vec{k}/\sqrt{\tau}, \vec{p}}} (1 - Q_{\frac{1}{\sqrt{\tau}}\vec{k},\vec{p}}) \widetilde{\rho}_{0;\frac{1}{\sqrt{\tau}} \vec{k}, \vec {p}}\otimes 1}. \label{2nd_term} \end{align}By Lemma \ref{lem:Lkdynamics}, the integrand in \eqref{2nd_term} is exponentially small in the large $\tau$ limit, \begin{multline}\label{eq:decay} \abs{\ip{ \delta_{0} \otimes 1, (1 - Q_{\frac{1}{\sqrt{\tau}}\vec{k} , \vec{p}}) \mathrm e^{-\tau t \widetilde{L}_{ \vec{k}/\sqrt{\tau}, \vec{p}}} \widetilde{\rho}_{0;\frac{1}{\sqrt{\tau}} \vec{k}, \vec {p}}\otimes 1}} \\ \le \ \norm{(1 - Q_{\frac{1}{\sqrt{\tau}}\vec{k} , \vec{p}}) \mathrm e^{-\tau t \widetilde{L}_{ \vec{k}/\sqrt{\tau}, \vec{p}}}} \norm{ \widetilde{\rho}_{0;\frac{1}{\sqrt{\tau}} \vec{k}, \vec {p}}\otimes 1} \ \le \ C_{\epsilon} \mathrm e^{-\tau t (\delta-\epsilon )} \ \rightarrow \ 0. \end{multline} Regarding \eqref{1st_term}, we have by Taylor's formula, \begin{equation} E_{\vec {p}}(\vec{k}/\sqrt{\tau}) \ = \ \frac{1}{2\tau} \sum_{i,j} \partial_{i} \partial_{j} E_{\vec {p}}(\vec 0) \vec{k}_{i} \vec{k}_{j} \ + \ o\left ( \frac{1}{\tau} \right ), \end{equation} since $E_{\vec {p}}(\vec 0) = \nabla E_{\vec {p}}(\vec 0)=0$. Thus \begin{equation}\label{eq:taylor} \mathrm e^{-\tau t E_{\vec {p}}(\vec{k}/\sqrt{\tau})} \ = \ \mathrm e^{-t \frac{1}{2} \sum_{i,j} \partial_{i} \partial_{j} E_{\vec {p}}(\vec 0) \vec{k}_{i} \vec{k}_{j}} + o(1), \end{equation} and \begin{multline}\label{eq:final} \sum_{x} \mathrm e^{-\mathrm i \frac{1}{\sqrt{\tau}} \vec{k} \cdot x} \Ev{\rho_{\tau t}(x,x)} \ = \ \frac{N^{d}}{(2\pi)^{d}} \int_{\bb{T}_{ N}^d}\mathrm d\vec{p} \ \mathrm e^{-t \frac{1}{2} \sum_{i,j} \partial_{i} \partial_{j} E_{\vec {p}}(\vec 0) \vec{k}_{i} \vec{k}_{j}} \ip{\delta_{0}\otimes 1, \widetilde{\rho}_{0;\frac{1}{\sqrt{\tau}} \vec{k}, \vec{p}}\otimes 1} + o(1) \\ \xrightarrow[]{\tau \rightarrow \infty} \ \frac{N^{d}}{(2\pi)^{d}} \int_{\bb{T}_{ N}^d}\mathrm d\vec{p}\ \mathrm e^{-t \frac{1}{2} \sum_{i,j} \partial_{i} \partial_{j} E_{\vec {p}}(\vec 0) \vec{k}_{i} \vec{k}_{j}} \widetilde{\rho}_{0;\vec 0, \vec p}(0) \end{multline} since $Q_{\vec{k}, \vec{p}}^{\dagger} \delta_{0} \otimes 1 \rightarrow \delta_{0} \otimes 1$ as $\vec{k} \rightarrow 0$ and $\widetilde{\rho}_{0;\vec{k}, \vec{p}}(0)$ is continuous as a function of $\vec{k}$. Letting $D_{i,j}(\vec{p})= \frac{1}{2} \partial_{i} \partial_{j} E_{\vec{p}}(\vec 0)$ and $m(\vec{p}) = \frac{N^{d}}{(2\pi)^{d}} \widetilde{\rho}_{0;\vec 0, \vec p}(0)$ gives \eqref{eq:main} and completes the proof.\qed \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace} \providecommand{\MR}{\relax\ifhmode\unskip\space\fi MR } \providecommand{\MRhref}[2]{% \href{http://www.ams.org/mathscinet-getitem?mr=#1}{#2} } \providecommand{\href}[2]{#2}
\section{Introduction} One main objective of the {\it INTEGRAL} hard X--ray satellite is the regular survey of the whole sky at high energies (above 20 keV). This makes use of the unique imaging capabilities of the IBIS instrument\cite{ibis} which permits the detection of sources at the mCrab level with a typical localization accuracy of 2-3 arcmin above 20 keV. A substantial fraction of the remaining objects ($\sim$30\%; see e.g. Ref.~\refcite{ajb}) had no obvious counterpart at other wavelengths and therefore cannot be associated with any known class of high-energy emitting sources. To this aim, since 2004 our group has been actively performing a successful observational campaign for the optical identification of these unidentified objects: up to now we identified more than 100 {\it INTEGRAL} sources and found that about half of them are Active Galactic Nuclei (AGNs), mostly located in the nearby Universe (at redshift $z \lesssim$ 0.1; see Ref.~\refcite{maso} and references therein). Here we present the continuation of this work with the use of the new {\it INTEGRAL} detections of unidentified sources reported in the recently published 4$^{\rm th}$ IBIS survey.\cite{ajb} \section{Sample selection} Using the criterion applied to our past works (see e.g. Ref.~\refcite{maso}), we positionally cross-correlated the 208 unidentified objects belonging to the 4$^{\rm th}$ IBIS survey\cite{ajb} with X--ray ({\it ROSAT}, {\it XMM-Newton}), radio (NVSS, SUMSS, MGPS) and far-infrared (IRAS) catalogues -- available online using SIMBAD\footnote{http://simbad.u-strasbg.fr} -- and with archival X--ray observations ({\it Swift}, {\it Chandra}). This was made in order to reduce the source error circle and pinpoint the putative optical counterpart. A few cases in which a catalogued emission-line optical object is present in the IBIS error box were also considered. We also included in our sample the source IGR J01054$-$7253, not included in the 4$^{\rm th}$ IBIS survey but recently discovered during an {\it INTEGRAL} key programme observation,\cite{boz} and for which an arcsec-sized X--ray position obtained with {\it Swift} is available in the literature.\cite{coe} This allowed us to pick out, and perform optical spectroscopy on, 25 candidate counterparts. The list of selected objects is reported in the 1st column of Table 1. \section{Caveats} While it is known\cite{jbs} that the presence of a single, bright soft X--ray object within the IBIS error circle indicates that it is with a high probability the lower-energy counterpart of the corresponding {\it INTEGRAL} source, the same cannot be said for radio and far-infrared sources, or for optically peculiar objects. Thus, for the IBIS sources which were selected using catalogues at wavelengths longer than soft X--rays, we caution the reader that the association with the selected optical object should await confirmation via a pointed observation with a soft X-ray satellite such as {\it XMM-Newtom}, {\it Chandra} or {\it Swift}. To stress this issue, when needed, we put an asterisk beside the source name in Table 1. \section{Observations} In this work, 6 telescopes located in Northern and Southern Hemisphere observatories were used; moreover, spectra from the 6dF\cite{6df} and SDSS\cite{sdss} on-line archives were also employed. In the following, the list of the telescopes used for the classification reported in Table 1 is given: \begin{itemlist} \item 1.5m telescope at CTIO, Chile; \item 1.5m "Cassini" telescope in Loiano, Italy; \item 1.8m "Copernico" telescope in Asiago, Italy; \item 1.9m "Radcliffe" telescope at SAAO, South Africa; \item 2.1m telescope in San Pedro M\'artir, Mexico; \item 2.5m telescope at Apache Point Observatory, New Mexico, USA, for the SDSS spectra; \item 3.58m NTT telescope at ESO-La Silla, Chile; \item 3.9m AAT telescope of the Anglo-Australian Observatory in Siding Spring, Australia, for the 6dF spectra. \end{itemlist} \section{Results} Table 1 reports, for each selected object, the corresponding classification obtained through optical spectroscopy, the object redshift and the name of the telescope used for the identification; Fig. 1 reports the spectra of the optical counterparts of some of these 25 selected IBIS sources. \begin{table \tbl{The sample of 25 {\it INTEGRAL} objects identified in this work together with their classification, redshift and telescope with which the identification was obtained. The asterisks indicate sources for which an X--ray position is still not available (see Sect.~3).} {\begin{tabular}{@{}llll@{}} \toprule Object name & Class & redshift & Telescope \\ \colrule IGR J00158+5605 & Sy1.5 & 0.168 & SPM \\ IGR J00465$-$4005 & Sy2 & 0.201 & CTIO \\ IGR J01054$-$7253 & HMXB & 0 & SAAO \\ IGR J01545+6437$^*$ & Sy2 & 0.034 & Asiago \\ IGR J02086$-$1742 & Sy1.2 & 0.129 & SPM \\ IGR J05253+6447 & likely Sy2 & 0.071 & Loiano \\ 1RXS J080114.6$-$462324 & CV & 0 & NTT \\ MCG +04$-$26$-$006 & LINER & 0.020 & SPM \\ IGR J1248.2$-$5828 & Sy1.9 & 0.028 & SAAO \\ IGR J13187+0322$^*$ & QSO & 0.606 & SDSS \\ IGR J14301$-$4156$^*$ & Sy2/LINER & 0.039 & 6dF \\ IGR J15311$-$3737 & Sy1 & 0.127 & SAAO \\ IGR J15549$-$3739$^*$ & Sy2 & 0.019 & 6dF \\ IGR J16287$-$5021 & LMXB & 0 & NTT \\ IGR J16327$-$4940$^*$ & HMXB & 0 & SAAO \\ 1RXS J165443.5$-$191620 & CV & 0 & SPM \\ IGR J18311$-$3337$^*$ & Sy2 & 0.066 & 6dF \\ IGR J19077$-$3925 & Sy1.9 & 0.073 & 6dF \\ IGR J19113+1533$^*$ & HMXB & 0 & SPM \\ IGR J19118$-$1707$^*$ & Sy2/LINER & 0.024 & 6dF \\ PKS 1916$-$300 & Sy1.5/1.8 & 0.167 & 6dF \\ IGR J19552+0044 & CV & 0 & Loiano \\ 1RXS J211336.1+542226 & CV & 0 & SPM \\ 1RXS J211928.4+333259 & Sy1.5/1.8 & 0.051 & SPM \\ 1RXS J213944.3+595016 & Sy1.5 & 0.114 & SPM \\ \botrule \end{tabular}} \end{table} \begin{figure \vspace{-2cm} \centerline{\psfig{file=fig1_hepro.ps,width=17cm}} \vspace{-3.5cm} \caption{Optical spectra of the counterparts of 8 objects belonging to the sample of 25 {\it INTEGRAL} sources identified in this work and listed in Table 1. These data allowed us to securely determine the nature and the redshift of these objects through inspection of their absorption and emission spectral features. The spectra are not corrected for the intervening Galactic absorption. For each spectrum the main spectral features are labeled. The symbol $\oplus$ indicates atmospheric telluric absorption bands.} \end{figure} As one can see from Table 1, most objects (17, i.e., 68\%) have an extragalactic origin, and are nearly equally divided into Type 1 and Type 2 AGNs (9 and 8 cases, respectively). Six of them (5 of Type 1 and only 1 of Type 2) have $z >$~0.1, indicating that this new, deeper IBIS survey is able to detect more distant AGNs. Only 32\% (8 sources) of our identifications are instead of Galactic nature. Interestingly, half of them are (likely magnetic) Cataclysmic Variables (CVs): these identifications increase by 13\% the number of hard X-ray emitting CVs detected with {\it INTEGRAL} up to now (see Ref.~\refcite{ss}). \section{Conclusions} We here presented the identification of a first set of 25 sources of unknown or uncertain nature, 24 of which belonging to the 4$^{\rm th}$ IBIS survey. We found that 2/3 of them have an extragalactic nature, with redshifts in the range 0.019--0.606, while the remaining ones are Galactic sources. It is noteworthy that the majority of the latter ones are (possibly magnetic) CVs. These preliminary results further confirm the {\it INTEGRAL} capabilities of detecting AGNs and hard X-ray emitting CVs. We also stress that with this work we already reduced by 12\% the number of unidentified sources of the 4$^{\rm th}$ IBIS survey, bringing it from 208 to 184. \section*{Acknowledgments} This research has made use of the SIMBAD database operated at CDS, Strasbourg, France, of the NASA/GSFC's HEASARC archive, and of the HyperLeda catalogue operated at the Observatoire de Lyon, France. We thank the referee for useful indications. The authors acknowledge the ASI and INAF financial support via grant No. I/008/07. PP is supported by the ASI-INTEGRAL grant No. I/008/07. VC is supported by the CONACYT research grant 54480-F (Mexico).
\section{Introduction} Chiral symmetry breaking is one the key feature of QCD. There is a very intuitive way of stating the physics of this phenomenon: a small quark mass leads to a macroscopic realignment of the QCD vacuum (this is a strict quotation from~\cite{JVTilo}). Since the QCD partition function reads \begin{equation}\label{eqn:Z} Z = \langle \prod_f \det ({\cal D} + m_f) \rangle = \langle \prod_f \prod_n (i \lambda_n + m_f) \rangle \end{equation} \noindent in order for this to be possible there must be an accumulation of Dirac operator eigenvalues near zero (otherwise the effect of a small quark mass would be overwhelmed by much larger eigenvalues). This message is actually encoded in the Banks Casher relation~\cite{BaCa} \begin{equation}\label{eqn:BC} \langle \bar{\psi} \psi \rangle = \frac{\pi \rho(0)}{V} \end{equation} \noindent relating the chiral condensate (the order parameter of the transition associated to spontaneous symmetry breaking) to the density of eigenvalues of the Dirac operator spectrum \begin{equation}\label{eqn:rho} \rho(\lambda) = \langle \sum_n \delta (\lambda - \lambda_n) \rangle. \end{equation} Altought not a natural observable in Field Theory, the Dirac operator spectrum has in force of~(\ref{eqn:BC}) become a natural probe for the chiral transition. Recent work~\cite{LeoLusch} has investigated the field theoretic status of spectral observables, in particular with respect to their renormalization properties. From a numerically point of view, it should be pointed out that Lattice QCD can quite naturally compute~(\ref{eqn:rho}), once a lattice regularization of the Dirac operator is given. \section{The Dirac spectrum and Perturbation Theory} Since the free Dirac operator has a vanishing eigenvalues density near zero, one is lead to the conclusion that the small eigenvalues are due to gauge interactions. There is actually a natural candidate: any quantum interaction produces a repulsion among the eigenvalues. With this respect Perturbation Theory is in a tantalizing situation: \begin{itemize} \item on one side, it sits (deep) in the chirally restored regime, while one looks for an effect which lives at its border; \item on the other side, it gives a unique opportunity to follow the fate of eigenvalues in their mutual repulsion. \end{itemize} We want to emphasize that our work is still at a very preliminary stage. In particular, we do not want to address here the subtleties which arise in properly defining a perturbative expansion of~(\ref{eqn:rho}). We will discuss a \emph{quick and dirty} procedure in which we first compute the perturbative corrections to the free spectrum eigenvalues \[ \lambda_n = \lambda_n^{(0)} + \beta^{-1/2} \lambda_n^{(1)} + \beta^{-1}\lambda_n^{(2)} + \ldots \] and then resum the expansion at given values of the coupling $\beta$. Given these summations, we can proceed to compute a density of eigenvalues much the same as in non-perturbative computations of the spectrum. \section{The Dirac Spectrum in NSPT} Numerical Stochastic Perturbation Theory~\cite{NSPT} relies on an expansion of the solution of Langevin equation. In the case of LGT \begin{equation}\label{eqn:NSPT} U_{x\mu}(\tau;\eta) \rightarrow 1+ \sum_{k=1} \beta^{-k/2} U^{(k)}_{x\mu}(\tau;\eta). \end{equation} \noindent $\tau$ is the stochastic time of Langevin evolution, with gaussian noise $\eta$. For asymptotic values of the stochastic time, $\eta$-averages $\langle \ldots \rangle_{\eta}$ of observables converge order by order to quantum field theory averages $\langle \ldots \rangle_{QFT}$.\\ Plugging~(\ref{eqn:NSPT}) into the Dirac operator turns the computation of~(\ref{eqn:rho}) into the typical eigenvalue/eigenvector problem in PT \begin{equation} M = M_0 + N = M_0 + \sum_i g^i N_i \;\;\;\;\;\;\;\;\; M \, |\alpha\rangle = \epsilon \, |\alpha\rangle \end{equation} which has to be solved by \begin{equation} \epsilon = \epsilon_0 + g \, \epsilon_1 + g^2 \, \epsilon_2 + \ldots \;\;\; |\alpha\rangle = |\alpha_0\rangle + g \, |\alpha_1\rangle + g^2 \, |\alpha_2\rangle + \ldots \end{equation} \noindent Due to the (huge) degeneracy of the free field solution, for every eigenvalue we need to explicitly separate components inside and outside the starting (degenerate) eigenspace, \emph{i.e.} \begin{equation} |\alpha\rangle = |\alpha_0\rangle + P'_{in} |\alpha\rangle + P_{out} |\alpha\rangle. \end{equation} \noindent In the previous formula $|\alpha_0\rangle$ is the direction in the free (degenerate) eigenspace singled out as the zeroth order of the solution; $P'_{in}$ is the projector onto the component of the free eigenspace which is orthogonal to $|\alpha_0\rangle$; $P_{out}$ projects instead outside the free eigenspace. We finally get the iterative solution \begin{eqnarray}\label{eqn:solution} \epsilon_n &=& \sum_{k=0}^{n} \, \langle \alpha_0 |N_{n-k}| \alpha_k\rangle \\ \nonumber P_{out} | \alpha \rangle &=& (\epsilon - M_0 - P_{out}N )^{-1} \left( \, P_{out}N |\alpha_0 \rangle + P_{out}N P'_{in} |\alpha \rangle \, \right)\\ \nonumber P'_{in} | \alpha \rangle &=& \; (\epsilon - \epsilon_0 - P'_{in}N )^{-1} \;\; \left( \, P'_{in}N |\alpha_0 \rangle + P'_{in}N P_{out}|\alpha \rangle \, \right).\\ \nonumber \end{eqnarray} This is the (closed) solution only provided degeneracy is lifted at first order. Should this not be the case, the formalism should be generalized by introducing a new projector for each level of degeneracy still present (the solution is nevertheless closed also in such a situation, which actually occurs in our computations). \\ In standard non-perturbative LGT computations of the Dirac operator spectrum one gets distributions of eigenvalues by generating configurations and computing the spectrum on each of them. The density of eigenvalues is then simply obtained by plain histograms of the results. We stress once again that at this stage of our work we will adhere to the \emph{naive} recipe of first computing the eigenvlaues in PT, then summing the expansions at given values of the coupling and finally constructing histograms much the same way as in the non-perturbative case. \section{Results} In figure~(1) we plot examples of our results: we collect all the measurements for first (trivial) and second (one loop) order corrections to free field results for the second lowest lying eigenspace on a $6^4$ lattice. We stress that this eigenspace is degenerate (the dimension of this eigenspace is $144$), but on top of this degeneracy the histograms entail the multeplicity which comes from the number of measurements. \begin{figure}[!htb] \begin{center} \includegraphics[scale=0.62,clip=true] {domeNbell.eps} \end{center} \vspace{-5mm} \caption{First (trivial) and second (one loop) corrections to the second (lowest lying) free field eigenvalue on a $6^4$ lattice (overall distributions of the measures).} \label{fig:} \end{figure} \begin{figure}[!htb] \begin{center} \includegraphics[scale=0.62,clip=true] {newE1E2.eps} \end{center} \vspace{-5mm} \caption{First (trivial) corrections to the first (lowest lying) and second free field eigenvalue on a $6^4$ lattice (averages over Langevin histories). Free eigenspace degeneracies are 24 (left) and 144 (right).} \label{fig:} \end{figure} Figure~(2) displays data once the average over all the measurements has been taken. In this case we plot first order corrections (as one expects, they average to zero) for lowest lying and second lowest lying eigenvalues. There are issues which are worth stressing. First of all, one can ispect degeneracies which are not lifted. Second, the distributions of corrections in the two eigenspaces differ quite a lot.\\ Figure~(3) displays another interesting feature. In this case we plot a third order correction, which enlights how higher orders display long tails. One probably needs to carefully assess when the free field degeneracy is actually lifted, as it is clear from the impact of denominators in~(\ref{eqn:solution}).\\ \begin{figure}[!htb] \begin{center} \includegraphics[scale=0.62,clip=true] {crazyg3.eps} \end{center} \vspace{-5mm} \caption{Third correction to the second free field eigenvalue on a $6^4$ lattice: while it is centered in zero (as expected), it displays long tails.} \label{fig:} \end{figure} With this respect we point out that one can always check the accuracy of the computation by considering quantities like \[ \;\;\;\;\;\;\;\; \langle \mbox{Tr} (D^\dagger D)^k \rangle = \ldots \;\;\;\;\;\;\;\; \langle \mbox{Tr} (D^\dagger D)^{-k} \rangle = \ldots \] They can be both computed directly and reconstructed from the eigenvalues distribution, eventually validating the latter. \\ We can now go back to our \emph{quick and dirty} procedure to inspect the impact of the perturbative corrections. Basically, we can sum the contributions at any given value of the coupling and try to follow the resulting distribution of eigenvalues as the gauge intercation comes into play. We plot in figure~(4) what we get at one loop. \begin{figure}[!htb] \begin{center} \includegraphics[width=\textwidth,height=12cm,clip=true] {moving.eps} \end{center} \vspace{-5mm} \caption{The evoultion of the eigenvalues density: from free field limit ($\beta=\infty$) to the intercating case (at different values of the coupling $\beta$).} \label{fig:} \end{figure} Figure~(4) is something like a sequence of pictures taken while the interaction is switched on. One starts at zero coupling, where the key feature of the free field is on display: bins are centered where free field eigenvalues sit, and bins heigth simply entails the degeneracy of the various eigenspaces. Notice anyway that at this resolution some bins actually results from the contribution of two free field eigenvalues sitting very close to each other. While the interaction is switched on (\emph{i.e.} the value of the inverse coupling $\beta$ decreases) the bins spread and overlap and eventually a non-zero density near zero is generated. A natural question arises: where do eigenvalues moving to zero come from? One should remember the point we made on repulsion among eigenvalues. Figure~(5) displays an example of how this takes place: we plot the contribution to $\rho$ coming from two eigenvalues starting very close to each other in free field.\\ It is worth better assessing the impact of the repulsion among the couple of free eigenvalues we have just looked at. It actually turns out that they give a substantial contribution to the rearrangement of the eigenvalues density: one can recognize their splitting on the right of figure~(6). \begin{figure}[!htb] \begin{center} \includegraphics[width=\textwidth,height=11.5cm,clip=true] {kicking_eo.eps} \end{center} \vspace{-5mm} \caption{Following the repulsion of two eigenvalues on $6^4$. They start very close in free field limit and then strongly repel each other.} \label{fig:} \end{figure} \begin{figure}[!hbt] \begin{center} \includegraphics[scale=0.62,clip=true] {kick6_5122.eps} \end{center} \vspace{-5mm} \caption{The first 5122 free eigenvalues on a $6^4$ lattice (black line): the lenghts of each segment is the degeneracy in free field. Red curve displays how they move at one loop at $\beta=7.5$.} \label{fig:} \end{figure} Black line in figure~(6) is nothing but another way of plotting the first row of figure~(4): we plot the first 5122 free eigenvalues and the lenght of each plateaux is just the degeneracy of each free field eigenspace. The superimposed red line shows the summation (at first loop) of the perturbative series for these eigenvalues at $\beta=7.5$.\\ Some caveats are of course in order: \begin{itemize} \item Is this a finite-volume effect? At the moment we have actually got the same qualitative picture at any (still moderate) size we studied. \item Is this a finite $a$ effect? Testing this is more difficult. \item One should carefully take care of the order of limits which is in place in the Banks Casher relation. \end{itemize} A few following steps are on their way: we will repeat the computation in the background of different $Z(3)$ vacua and to try to reconstruct the Polyakov loop from the spectral decomposition of the Dirac operator (this is in the spirit of recent works by Gattringer~\cite{CGatt}). \section{Conclusions} Even though at a very preliminary stage, we showed some results of a perturbative computation of the Dirac operator spectrum by means of NSPT. Our results quantitatively support the picture of the repulsion among eigenvalues being responsible for the rearrangement of eigenvalues, ultimately giving rise to Banks Casher.\\ Some developments of this work are expected to follow these preliminary results. \begin{itemize} \item We have to carefully assess the huge tails of higher order distributions. One hint is that this asks for some regulator in highly degenerate free field eigenspaces. \item We will move to computations in the background of different $Z(3)$ vacua. \item Having at hand full spectra could in principle enable the computation of a variety of quantities. \end{itemize}
\section{Introduction} \emph{A preliminary announcement (without proofs) of the results in this paper is to appear in \emph{Electronic Research Announcements}.} \subsection{Overview of multifractal formalism} The basic elements of the multifractal formalism were first proposed by Halsey \emph{et al} in~\cite{HJKPS86}, where they considered what they referred to as the \emph{dimension spectrum} or the \emph{$f(\alpha)$-spectrum for dimensions}, which characterises an invariant measure $\mu$ for a dynamical system $f\colon X\to X$ in terms of the level sets of the \emph{pointwise dimension}. The pointwise dimension of $\mu$ at $x$ is defined as \[ d_\mu(x) = \lim_{\eps\to 0} \frac{\log \mu(B(x,\eps))}{\log\eps}, \] provided the limit exists, and the level sets are denoted \[ \Kd{\alpha} = \{ x\in X \mid d_\mu(x) = \alpha \}. \] Many measures of interest are \emph{exact-dimensional}; that is, the pointwise dimension is constant $\mu$-almost everywhere. In particular, this is true of hyperbolic measures (those with non-zero Lyapunov exponents almost everywhere)~\cite{BPS99}. For an exact-dimensional measure, one of the $\Kd{\alpha}$ has full measure, and the rest have measure $0$, and so we measure the sizes of these sets with the Hausdorff dimension rather than with the measure; in this way we obtain the \emph{dimension spectrum for pointwise dimensions}, which is given by the function \[ \DDD(\alpha)=\dim_H \Kd{\alpha}. \] One goal of the multifractal formalism is to show that under certain conditions on $f$ and $\mu$, the function $\mathcal{D}$ is in fact analytic and concave on its domain of definition, and is related to the R\'enyi and Hentschel--Procaccia spectra for dimensions by a Legendre transform. This was done by Rand~\cite{dR89} when $\mu$ is a Gibbs measure on a hyperbolic cookie-cutter (a dynamically defined Cantor set), and by Pesin and Weiss~\cite{PW97} for uniformly hyperbolic conformal maps: modern expositions of the whole theory for uniformly hyperbolic systems can be found in~\cite{yP98,BPS97,TV00}. More recently, various non-uniformly hyperbolic systems have been studied in~\cite{kN00,mT08,JR09,IT09b}. There are other important examples of multifractal spectra; each such spectrum measures the level sets of some local quantity by using a global (dimensional) quantity. For $\DDD(\alpha)$, these roles are played by pointwise dimension and Hausdorff dimensions, respectively; one may also consider spectra defined using other quantities. For example, one may consider the measure of small balls which are refined dynamically, rather than statically. That is, rather than $B(x,\eps)$ we consider the Bowen ball of radius $\delta$ and length $n$, given by \[ B(x,n,\delta) = \{y\in X\mid f^k(y)\in B(f^k(x),\delta) \text{ for } k=0,1,\dots,n \}. \] If the map $f$ has some eventual expansion, then the balls $B(x,n,\delta)$ decrease in size, and in measure, as $n$ increases with $\delta$ held fixed. Just as the rate at which $\mu(B(x,\eps))$ decreases with $\eps$ is the pointwise dimension $d_\mu(x)$, so also the rate at which $\mu(B(x,n,\delta))$ decreases with $n$ is the \emph{local entropy} of $\mu$ at $x$ \[ h_\mu(x) = \lim_{\delta\to 0} \llim_{n\to\infty} -\frac{1}{n} \log \mu(B(x,n,\delta)), \] provided the limit exists. We denote the level sets of the local entropy by \[ \Ke{\alpha} = \{x\in X \mid h_\mu(x) = \alpha \}. \] It was shown by Brin and Katok that if $\mu$ is ergodic, then one of the level sets $\Ke{\alpha}$ has full measure, and the rest have measure $0$~\cite{BK83}; thus we must once again quantify them using a (global) dimensional characteristic. It turns out to be more natural to measure the size of the sets $\Ke{\alpha}$ with the topological entropy rather than Hausdorff dimension; because these level sets are in general not compact, we must use the definition of topological entropy in the sense of Bowen~\cite{rB73}. Upon doing so, we obtain the \emph{entropy spectrum for local entropies} \[ \EEE(\alpha) = h_\mathrm{top} (\Ke{\alpha}). \] For Gibbs measures on conformal repellers, this spectrum was studied in~\cite{BPS97}. Takens and Verbitskiy~\cite{TV99} carred out the multifractal analysis in the more general case of expansive maps satisfying a specification property. The Gibbs property of the measures studied so far is essential, because it relates local scaling quantities of the measure (pointwise dimension or local entropy) to asymptotic statistical properties of a potential function $\ph$. In fact, the proofs of the known results for both the dimension and entropy spectra contain (at least implicitly) a similar result for the \emph{Birkhoff spectrum}. Writing the sum of $\ph$ along an orbit as $S_n\ph(x) = \sum_{k=0}^{n-1} \ph(f^k(x))$, the \emph{Birkhoff average} of $\ph$ at $x$ is given by \[ \ph^+(x) = \lim_{n\to\infty} \frac{1}{n} S_n \ph(x), \] provided the limit exists. The level sets of the Birkhoff averages are \[ \Kb{\alpha} = \{x\in X \mid \ph^+(x) = \alpha \}, \] and the Birkhoff ergodic theorem guarantees that for any ergodic measure $\mu$, one of the level sets has full measure, and the rest have measure $0$. Thus we once again measure their size in terms of topological entropy, and obtain the \emph{entropy spectrum of Birkhoff averages} \[ \BBB(\alpha) = h_\mathrm{top} (\Kb{\alpha}). \] In the uniformly hyperbolic setting, results on the Birkhoff spectrum were obtained in~\cite{PW01}, among other places. More general results, some of which overlap with one of the results in this paper, were recently announced in~\cite{FH10}. The general scheme tying all these spectra together is as follows. Given an asymptotic local quantity---pointwise dimension, local entropy, Birkhoff average---we have an associated multifractal decomposition into level sets of this quantity. These level sets are then measured using a global dimensional quantity---Hausdorff dimension or topological entropy. This defines a multifractal spectrum, which associates to each real number $\alpha$ the dimension of the level set corresponding to $\alpha$. Following this general outline, each of the above spectra could also be defined using the alternate global dimensional quantity. That is, we could define the \emph{entropy spectrum for pointwise dimensions} by \[ \mathcal{D}_E(\alpha) = h_\mathrm{top}(\Kd{\alpha}), \] and similarly for the \emph{dimension spectrum for local entropies} and the \emph{dimension spectrum for Birkhoff averages}. It turns out that these \emph{mixed multifractal spectra} are harder to deal with than the ones we have defined so far; see~\cite{BS01} for further details. We will restrict our attention to the spectra for which the local and global quantities are naturally related, and will generally simply refer to the \emph{entropy spectrum}, the \emph{dimension spectrum}, and the \emph{Birkhoff spectrum}. We will see that the Birkhoff spectrum provides a simpler setting for arguments which also apply to the dimension and entropy spectra; it is also of interest in its own right, having applications to the theory of large deviations~\cite{PW01, BR87}. One important example of a Birkhoff spectrum is worth noting. In the particular case where $f$ is a conformal map and $\ph(x)=\log \|Df(x)\|$, the Birkhoff averages coincide with the Lyapunov exponents: $\lambda(x) = \ph^+(x)$. In this case we will also denote the level sets by \[ \Kl{\alpha} = \{ x\in X \mid \lambda(x) = \alpha \}; \] it turns out that we are able to examine not only the \emph{entropy spectrum for Lyapunov exponents} \[ \LLL_E(\alpha) = h_\mathrm{top} \Kl{\alpha}, \] but also the \emph{dimension spectrum for Lyapunov exponents} \[ \LLL_D(\alpha) = \dim_H \Kl{\alpha}, \] by using a generalisation of Bowen's equation to non-compact sets~\cite{BS00,vC09a}. We may refer to either $\LLL_E(\alpha)$ or $\LLL_D(\alpha)$ as the \emph{Lyapunov spectrum}. It is often the case that the ``interesting'' dynamics takes place on a repeller which has Lebesgue measure zero---the Lyapunov spectrum provides information on how quickly the trajectories of nearby points escape from a neighbourhood of the repeller~\cite{BR87}. Taken together, the various multifractal spectra provide a great deal of information about the map $f$. In fact, certain classes of systems are known to exhibit \emph{multifractal rigidity}, in which a finite number of multifractal spectra completely characterise a map~\cite{BPS97}. \subsection{General description of results} Direct computation (numerical or otherwise) of the various multifractal spectra is quite difficult. In the first place, in order to determine the level sets $K_\alpha$, one needs to first compute the asymptotic quantity (Birkhoff average, pointwise dimension, local entropy) at \emph{every} point of $X$. Even if this is accomplished, it still remains to compute the (Bowen) topological entropy or Hausdorff dimension of $K_\alpha$ for every value of $\alpha$. Because this quantity is defined as a critical point, rather than as a growth rate, it is more difficult to compute than the (capacity) topological entropy or the box dimension. (These latter quantities are of little use in analysing the level sets $K_\alpha$ since they assign the same value to a set and to its closure, and the level sets $K_\alpha$ are dense in many natural situations.) Rather than a direct frontal assault, then, the most successful method for analysing multifractal spectra has been to relate them to certain thermodynamic functions via the Legendre transform. These thermodynamic functions, which are given in terms of the topological pressure, can be computed more easily than the multifractal spectra, as they are given in terms of the growth rates of a family of partition functions. This approach goes back to~\cite{dR89} (the Legendre transform appeared already in~\cite{HJKPS86}, but in terms of the Hentschel--Procaccia and R\'enyi spectra, not in terms of the topological pressure). To date, the general strategy informed by this philosophy has been as follows: \begin{enumerate}[(1)] \item Fix a specific class of systems---uniformly hyperbolic maps, conformal repellers, parabolic rational maps, Manneville--Pomeau maps, multimodal interval maps, etc. \item Using tools specific to that class of systems (Markov partitions, specification, inducing schemes), establish thermodynamic results---existence and uniqueness of equilibrium states, differentiability of the pressure function, etc. \item Using these thermodynamic results \emph{together with the original toolkit}, study the multifractal spectra, and show that they can be given in terms of the Legendre transform of various pressure functions. \end{enumerate} Despite the success of this approach for a number of different classes of systems, there do not appear to be any extant rigorous results which apply to general continuous maps and arbitrary potentials (but see the remark below concerning~\cite{FH10}). Such results would give information about the multifractal analysis in settings far beyond those already considered; they would also establish the multifractal analysis as a direct corollary of the thermodynamic formalism, rendering Step (3) above automatic, and eliminating the need for the use of a specific toolkit to study the multifractal formalism itself. The results of this paper are a step in this direction. Not only do we obtain results that apply to general continuous maps regarding which nothing had been known, but the results described below also give alternate proofs of most previously known multifractal results, which are in some cases more direct than the original proofs. We obtain our strongest result for the Birkhoff spectrum $\BBB(\alpha)$. This result is given in Theorem~\ref{thm:birkhoff}, which applies to continuous maps $f\colon X\to X$ and to functions $\ph\colon X\to \RR$ which lie in a certain class $\AAA_f$; this class contains, but is not limited to, the space of all continuous functions. For such maps and functions, we show that the function $T_\BBB\colon q\mapsto P(q\ph)$, where $P$ is the pressure, is the Legendre transform of $\BBB(\alpha)$, \textbf{\emph{without any further restrictions on $f$ and $\ph$}}. Furthermore, we show that $\BBB(\alpha)$ is the Legendre transform of $T_\BBB$, completing the multifractal formalism, \textbf{\emph{provided $T_\BBB$ is continuously differentiable and equilibrium measures exist}}. If the hypotheses on $T_\BBB$ only hold for certain values of $q$, we still obtain a partial result on $\BBB(\alpha)$ for the corresponding values of $\alpha$. \begin{remark} After this paper was completed, the author was made aware of recent results announced by Feng and Huang in~\cite{FH10}, which deal with asymptotically sub-additive sequences of potentials, and which include Theorem~\ref{thm:birkhoff} for continuous potentials $\ph$ as a special case (however, they do not consider any of the dimension spectra). Many of the methods of proof are similar, and it appears as though the other results in this paper could also be extended to the non-additive case they consider. We observe that due to their definition of pressure, which only applies to functions $\ph$ such that $e^{\ph(x)}$ is continuous, their results do not apply to the discontinuous potentials in $\AAA_f$, nor to the more general class of bounded measurable potentials for which we obtain partial results (see below). To the best of the author's knowledge, the present results are the first rigorous multifractal results for general discontinuous potentials. \end{remark} Theorem~\ref{thm:birkhoff} gives an alternate (and more direct) proof of the multifractal formalism for the Birkhoff spectrum of a H\"older continuous potential function and a uniformly hyperbolic system, which was first established by Pesin and Weiss~\cite{PW01}. It can also be applied to non-uniformly hyperbolic systems; in addition to some systems that have already been studied, we describe in Section~\ref{sec:app} a class of systems studied by Varandas and Viana~\cite{VV08} to which Theorem~\ref{thm:birkhoff} can be applied. Proposition~\ref{prop:VV} gives multifractal results for these systems; these results appear to be completely new. As stated, Theorem~\ref{thm:birkhoff} does not deal with phase transitions---that is, points at which the pressure function is non-differentiable. Such points correspond (via the Legendre transform) to intervals over which the Birkhoff spectrum is affine (if the multifractal formalism holds). In Theorem~\ref{thm:phase}, we give slightly stronger conditions on the map $f$, which are still fundamentally thermodynamic in nature, under which we can establish the complete multifractal formalism even in the presence of phase transitions. It is often the case that thermodynamic considerations demonstrate the existence of a \emph{unique} equilibrium state for certain potentials. In Proposition~\ref{prop:unique-works}, we show that if the entropy function is upper semi-continuous, then uniqueness of the equilibrium state implies differentiability of the pressure function and allows us to apply Theorem~\ref{thm:birkhoff}. However, Example~\ref{eg:vw} shows that there are systems for which the pressure function is differentiable, and hence Theorem~\ref{thm:birkhoff} can be applied, even though the equilibrium state is non-unique. One would like to understand for which classes of discontinuous potentials the multifractal formalism holds. Things work well for $\ph\in \AAA_f$ because the weak* topology is the same at $f$-invariant measures whether we consider continuous test functions or test functions in $\AAA_f$. Beyond this class of potentials, things are more delicate. We consider general measurable potentials that are bounded above and below, and while we do not obtain results for all values of $\alpha$, we do obtain in Theorem~\ref{thm:high-entropy} complete results for those values of $\alpha$ at which $\Tb^{L_1}$ is larger than the topological entropy of the closure of the set of discontinuities of $\ph$, and for the corresponding values of $q$. Ideally, we would be able to include \emph{unbounded} potentials in these results. In particular, we would like to be able to consider the geometric potential $\ph(x) = -\log |f'(x)|$ for a multimodal map $f$; the presence of critical points leads to singularities of $\ph$, and so $\ph$ is not bounded above. Theorem~\ref{thm:singularity} shows that the results of Theorem~\ref{thm:birkhoff} still hold for $q\leq 0$ (that is, values of $q$ such that $q\ph$ is bounded above) and for the corresponding values of $\alpha$. The question of what happens for $q>0$ is more delicate and remains open. In Section~\ref{sec:conformal}, we use a non-uniform version of Bowen's equation~\cite{vC09a} to give a result for the Lyapunov spectrum $\LLL_D(\alpha)$ in the case where $f$ is a conformal map without critical points, which satisfies some asymptotic expansivity properties. In order to obtain results on the spectra $\EEE(\alpha)$ and $\DDD(\alpha)$, for which the corresponding local quantities ($d_\mu(x)$ and $h_\mu(x)$) are defined in terms of an invariant measure $\mu$, we need some relationship between $\mu$ and a potential function $\ph$. This is given by the assumption that $\mu$ is a \emph{weak Gibbs measure} for $\ph$; we observe that there are several cases in which weak Gibbs measures (of one definition or another) are known to exist~\cite{mY00,mK01,FO03,VV08,JR09}. For such measures, we will see that the level sets $\Ke{\alpha}$ are determined by the level sets $\Kb{\alpha}$, and hence we obtain Theorem~\ref{thm:entropy}, which gives the corresponding result for the entropy spectrum $\EEE(\alpha)$ of a Gibbs measure, and follows from Theorem~\ref{thm:birkhoff}. Writing $\ph_1=\ph-P(\ph)$, we find $\EEE(\alpha)$ as the Legendre transform of the function $T_\EEE\colon q\mapsto P(-q\ph_1)$, \textbf{\emph{provided $T_\EEE$ is continuously differentiable and equilibrium measures exist}}. Theorem~\ref{thm:dimension} deals with the dimension spectrum $\DDD(\alpha)$ in the case where $f$ is conformal without critical points and $\mu$ is a weak Gibbs measure for a continuous potential $\ph$. Passing to $\ph_1$ so that $P(\ph_1)=0$, we follow Pesin and Weiss~\cite{PW97}, and define a family of potential functions $\ph_q$ by \begin{equation}\label{eqn:implicit} \ph_q(x) = -T_\DDD(q) \log \|Df(x)\| + q\ph_1(x), \end{equation} with $T_\DDD(q)$ chosen so that $P(\ph_q)=0$. Under mild expansivity conditions on $f$, we show that the implicitly defined function $T_\DDD(q)$ is the Legendre transform of the dimension spectrum $\DDD(\alpha)$, \textbf{\emph{without any further conditions on $f$ or $\ph$}}. Furthermore, we show that $\DDD(\alpha)$ is the Legendre transform of $T_\DDD(q)$, completing the multifractal formalism, \textbf{\emph{provided $T_\DDD$ is continuously differentiable and equilibrium measures exist}}. Results for all of the above spectra have already been known in specific cases. However, the present results differ from previous work in that their proofs do not use properties of the map $f$ that are specific to a particular class, but rather rely on thermodynamic results. This is particularly true of Theorem~\ref{thm:birkhoff}, which requires nothing at all of $f$ besides continuity. We also observe that the requirement of conformality in~\eqref{eqn:dimlyap} and Theorem~\ref{thm:dimension} is somehow unavoidable if we wish to use the any of the standard definitions of pressure; for a non-conformal map, one would need to consider a non-additive version of the pressure~\cite{lB96,FH10}, and it is not clear what implicit definition for $T_\DDD$ should replace~\eqref{eqn:implicit}. In Sections~\ref{sec:rmk} and~\ref{sec:app}, we make various general remarks concerning the results and their applications to both known and new examples. Sections~\ref{sec:prep} through~\ref{sec:last} contain the proofs. \emph{Acknowledgements.} Many thanks are due to my advisor, Yakov Pesin, for the initial suggestion to pursue this approach, and for much guidance and encouragement along the way. I would also like to thank Van Cyr, Katrin Gelfert, Stefano Luzzatto, Omri Sarig, Sam Senti, and Mike Todd for helpful conversations as this work took on its present form. \section{Definitions and results for Birkhoff spectrum} Throughout this section, we fix a compact metric space $X$, a continuous map $f\colon X\to X$, and a Borel measurable potential function $\ph\colon X\to \RR$. To fix notation, we recall the definition of Hausdorff dimension. \begin{definition} Given $Z\subset X$ and $\eps>0$, let $\mathcal{D}(Z,\eps)$ denote the collection of countable open covers $\{ U_i \}_{i=1}^\infty$ of $Z$ for which $\diam U_i\leq \eps$ for all $i$. For each $s\geq 0$, consider the set functions \begin{align} \label{eqn:mHse} m_H(Z,s,\eps) &= \inf_{\mathcal{D}(Z,\eps)} \sum_{U_i} (\diam U_i)^s, \\ \label{eqn:mHs} m_H(Z,s) &= \lim_{\eps\to 0} m_H(Z,s,\eps). \end{align} The \emph{Hausdorff dimension} of $Z$ is \[ \dim_H Z = \inf \{ s>0 \mid m_H(Z,s)=0 \} = \sup \{ s>0 \mid m_H(Z,s)=\infty \}. \] It is straightforward to show that $m_H(Z,s)=\infty$ for all $s<\dim_H Z$, and that $m_H(Z,s)=0$ for all $s>\dim_H Z$. \end{definition} An analogous definition of topological entropy was given by Bowen~\cite{rB73}, establishing it as another dimensional characteristic. \begin{definition} Given $Z\subset X$, $\delta>0$, and $N\in\NN$, let $\mathcal{P}(Z,N,\delta)$ denote the collection of countable sets $\{ (x_i,n_i) \}_{i=1}^\infty \subset Z\times\NN$ such that $\{B(x_i,n_i,\delta)\}$ covers $Z$ and $n_i\geq N$ for all $i$. For each $s\in\RR$, consider the set functions \begin{align} \label{eqn:mhNd} m_h(Z,s,N,\delta) &= \inf_{\mathcal{P}(Z,N,\delta)} \sum_{(x_i,n_i)} e^{-n_i s}, \\ \label{eqn:mhd} m_h(Z,s,\delta) &= \lim_{N\to\infty} m_h(Z,s,N,\delta), \end{align} and put \[ h_\mathrm{top}(Z,\delta) = \inf \{ s>0 \mid m_h(Z,s,\delta) = 0 \} = \sup \{ s>0 \mid m_h(Z,s,\delta) = \infty \}. \] As with Hausdorff dimension, we get $m_h(Z)=\infty$ for $s<h_\mathrm{top}(Z,\delta)$, and $m_h(Z)=0$ for $s>h_\mathrm{top}(Z,\delta)$. The \emph{topological entropy} of $f$ on $Z$ is \[ h_\mathrm{top}(Z) = \lim_{\delta\to 0} h_\mathrm{top}(Z,\delta). \] \end{definition} If we replace the quantity $e^{-n_i s}$ in~\eqref{eqn:mhNd} with $e^{n_i s + S_{n_i} \ph(x_i)}$, the definition above gives us not the topological entropy but the topological \emph{pressure} $P_Z(\ph)$, introduced in this form by Pesin and Pitskel' in~\cite{PP84} (although the version of pressure we will discuss below dates back to Ruelle~\cite{dR73} and Bowen~\cite{rB75b}). All three of these quantities (Hausdorff dimension, entropy, and pressure) are defined as critical points and have certain important properties common to a broad class of Carath\'eodory dimension characteristics (see~\cite{yP98} for details). We will use two of these repeatedly, so they are worth mentioning here: in the first place, given any countable family of sets $Z_i\subset X$, we have \[ \dim_H \left(\bigcup_i Z_i\right) = \sup_i \dim_H Z_i, \] and similarly for $h_\mathrm{top} Z$ and $P_Z(\ph)$. Furthermore, all of these quantities can be bounded above in terms of a corresponding \emph{capacity}; for Hausdorff dimension, the corresponding capacity is the lower box dimension. We recall the definitions of the analogues for entropy and pressure (see~\cite{yP98}). \begin{definition} A set $E\subset Z$ is $(n,\delta)$-spanning if $Z \subset \bigcup_{x\in E} B(x,n,\delta)$. The \emph{(lower) capacity topological entropy} $\underline{Ch}_\mathrm{top}(Z)$ is the lower asymptotic growth rate of the minimal cardinality of an $(n,\delta)$-spanning set in $Z$. More precisely, if $P_n^\delta$ is the minimal cardinality of such a set, then \begin{align} \label{eqn:lhZd} \underline{Ch}_\mathrm{top}(Z,\delta) &= \llim_{n\to\infty} \frac 1n \log P_n^\delta, \\ \label{eqn:lhZ} \underline{Ch}_\mathrm{top}(Z) &= \lim_{\delta\to 0} \underline{Ch}_\mathrm{top}(Z,\delta). \end{align} A similar definition taking the upper limit gives us $\overline{Ch}_\mathrm{top}(Z)$. In the proof of Theorem~\ref{thm:dimension}, we will also need the notion of \emph{capacity topological pressure}, whose definition we recall here. Fix a potential $\ph\colon X\to \RR$ and a subset $Z\subset X$. For every $n\in \NN$, $\delta>0$, let $E_n$ be a minimal $(n,\delta)$-spanning set: then the lower capacity topological pressure of $\ph$ on $Z$ is given by \begin{align} \label{eqn:lPZd} \underline{CP}_Z(\ph,\delta) &= \llim_{n\to\infty} \frac 1n \sum_{x\in E_n} e^{S_n \ph(x)}, \\ \label{eqn:lPZ} \underline{CP}_Z(\ph) &= \lim_{\delta\to 0} \underline{CP}_Z(\ph,\delta). \end{align} We have a corresponding definition of $\overline{CP}_Z(\ph)$. In the case $\ph=0$, these reduce to $\underline{Ch}_\mathrm{top}(Z)$ and $\overline{Ch}_\mathrm{top}(Z)$, respectively. Elementary arguments given in~\cite{pW75} show that we can also use maximal $(n,\delta)$-separated sets in the above definitions, and we will occasionally do so. \end{definition} We observe that the definitions given above differ slightly from the definitions in~\cite{yP98}. For a proof that both sets of definitions yield the same quantities when the potential $\ph$ is continuous, see~\cite[Proposition~4.1]{vC09a}. In general, we have the following relationship between the three pressures~\cite[(11.9)]{yP98}: \begin{equation}\label{eqn:pressures} P_Z(\ph) \leq \underline{CP}_Z(\ph) \leq \overline{CP}_Z(\ph). \end{equation} If $Z$ is compact and $f$-invariant (for example, if $Z=X$), then we have equality in~\eqref{eqn:pressures}, and the variational principle relates the common quantity to the following definition, which we will use to state our thermodynamic requirements. \begin{definition} Let $\MMM(X)$ be the set of all Borel probability measures on $X$, and denote by $\MMM^f(X)$ the set of $f$-invariant measures in $\MMM(X)$. Given $\mu\in\MMM^f(X)$, write $h(\mu)$ for the measure theoretic entropy of $\mu$. The \emph{(variational) pressure} of $\ph$ is \begin{equation}\label{eqn:pressure} P^*(\ph) = \sup \left\{ h(\mu) + \int \ph\,d\mu \,\Big|\, \mu\in \MMM^f(X) \right\} \end{equation} If $Z\subset X$ is compact and $f$-invariant, we will write the pressure on $Z$ as \[ P_Z^*(\ph) = \sup \left\{ h(\mu) + \int \ph\,d\mu \,\Big|\, \mu\in \MMM^f(Z) \right\}, \] where $\MMM^f(Z) = \{ \mu \in \MMM^f(X) \mid \mu(Z) = 1 \}$. Let $\MMM^f_E(X)$ be the set of all ergodic measures in $\MMM^f(X)$. It follows using the ergodic decomposition that~\eqref{eqn:pressure} is equivalent to \[ P^*(\ph) = \sup \left\{ h(\mu) + \int\ph\,d\mu \,\Big|\, \mu\in \MMM^f_E(X) \right\}. \] A measure $\nu\in\MMM^f(X)$ is an \emph{equilibrium state} for the potential $\ph$ if it achieves this supremum; that is, if \[ P^*(\ph) = h(\nu) + \int\ph\,d\nu. \] Every equilibrium state is a convex combination of ergodic equilibrium states. \end{definition} As is customary in multifractal formalism, we use the Legendre transform in the following slightly non-standard form. \begin{definition} Recall that a function $T\colon \RR\to [-\infty, +\infty]$ is convex if \begin{equation}\label{eqn:cvx} T(aq + (1-a)q') \leq aT(q) + (1-a) T(q') \end{equation} for all $0\leq a\leq 1$ and $q,q'\in \RR$. Given a convex function $T$, the \emph{Legendre transform} of $T$ is \begin{equation}\label{eqn:TL} T^{L_1}(\alpha) = \inf_{q\in \RR} (T(q) - q\alpha). \end{equation} Given a concave function $S\colon \RR\to [-\infty,+\infty]$ (for which the inequality in~\eqref{eqn:cvx} is reversed), the Legendre transform of $S$ is \begin{equation}\label{eqn:SL} S^{L_2}(q) = \sup_{\alpha\in \RR} (S(\alpha) + q\alpha). \end{equation} \end{definition} The Legendre transform of a convex function is concave, and vice versa. Furthermore, the Legendre transform is self-dual: if $T$ is convex and $T^{L_1} = S$, then $S^{L_2} = T$. Similarly, if $S$ is concave and $S^{L_2} = T$, then $T^{L_1} = S$. In what follows, we will consider situations in which the function $T$ is known to be convex (being given in terms of the pressure function), but the function $S$ is one of the multifractal spectra, about which we have no \emph{a priori} knowledge. Observe that the Legendre transform of such a function $S$ can still be defined by~\eqref{eqn:SL}, but in this case we lose duality; in its place we get the statement that $(S^{L_2})^{L_1}$ is the concave hull of $S$, the smallest concave function bounded below by $S$. Observe also that if $S(x)\geq 0$ for every $x\in \RR$, then $S^{L_2}$ is infinite everywhere. Thus for purposes of defining the various multifractal spectra, we adopt the (non-standard) convention that $h_\mathrm{top} \emptyset = \dim_H \emptyset = -\infty$. We recall that if $T$ is known to be convex, then left and right derivatives exist at every point where $T$ is finite; we will denote these by \[ D^- T(q) = \lim_{q'\to q^-} \frac{T(q) - T(q')}{q-q'}, \qquad D^+ T(q) = \lim_{q'\to q^+} \frac{T(q') - T(q)}{q'-q}. \] Existence follows from monotonicity of the slopes of the secant lines. Given a convex function $T$, define a map from $\RR$ to closed intervals in $\RR$ by $A(q) = [D^- T(q), D^+ T(q)]$. Extend this in the natural way to a map from subsets of $\RR$ to subsets of $\RR$; we will again denote this map by $A$. This map has the following useful property: given any set $I_Q\subset \RR$ and $\alpha\in A(I_Q)$, we have \[ T^{L_1}(\alpha) = \sup_{q\in I_Q} (T(\alpha) + q\alpha). \] This will be important for us in settings where we only have partial information about the functions $T$ and $S$. We will also make use of a map in the other direction: given a set $I_A \subset \RR$ (in the domain of $S$), we denote the set of corresponding values of $q$ by \[ Q(I_A) = \{q\in \RR \mid A(q)\cap I_A \neq \emptyset \}. \] In particular, if $\alpha = T'(q)$, then $\alpha = A(q)$, and if $q=-S'(\alpha)$, then $q = Q(\alpha)$. If $(q_1,q_2)$ is an interval on which $T$ is affine, then $A((q_1,q_2))$ is the slope of $T$ on that interval; furthermore, $T^{L_1}$ has a point of non-differentiability at $A((q_1,q_2))$. In the results below, it will sometimes be important to know whether or not $T$ is differentiable. A standard cardinality argument shows that $D^-T(q) = D^+T(q)$ at all but countably many values of $q$; however, the values of $q$ at which differentiability fails may \emph{a priori} be dense in $\RR$. Our most general result gives the following function as the Legendre transform of the Birkhoff spectrum: \begin{equation}\label{eqn:Tb} T_\BBB(q) = P^*(q\ph), \end{equation} Note that the function $T_\BBB$ is convex; even before we establish that $T_\BBB$ is the Legendre transform of $\BBB(\alpha)$, convexity follows immediately from the definition of variational pressure as a supremum and the fact that for every $\mu\in \MMM^f(X)$, the function $q\mapsto h(\mu) + \int q\ph\,d\mu$ is linear. Finally, before stating the general result, we describe the class of functions to which it applies. Given a function $\ph\colon X\to \RR$, let $\CCC(\ph)\subset X$ denote the set of points at which $\ph$ is discontinuous. Then we let $\AAA_f$ denote the class of Borel measurable functions $\ph\colon X\to \RR$ which satisfy the following conditions: \begin{enumerate}[(A)] \item $\ph$ is bounded (both above and below); \item $\mu(\overline{\CCC(\ph)}) = 0$ for all $\mu\in \MMM^f(X)$. \end{enumerate} In particular, $\AAA_f$ includes all continuous functions $\ph\in \CCC(X,\RR)$. It also includes all bounded measurable functions $\ph$ for which $\CCC(\ph)$ is finite and contains no periodic points, and more generally, all bounded measurable functions for which $\overline{\CCC(\ph)}$ is disjoint from all its iterates. We will see later (Proposition~\ref{prop:convergence}) that passing from $\CCC(X,\RR)$ to $\AAA_f$ does not change the weak* topology at measures in $\MMM^f(X)$, which is the key to including these particular discontinuous functions in our results. \begin{theorem}[The entropy spectrum for Birkhoff averages]\label{thm:birkhoff} Let $X$ be a compact metric space, $f\colon X\to X$ be continuous, and $\ph\in \AAA_f$. Then \begin{enumerate}[I.] \item $T_\BBB$ is the Legendre transform of the Birkhoff spectrum: \begin{equation}\label{eqn:TisBL} T_\BBB(q) = \BBB^{L_2}(q) = \sup_{\alpha\in\RR} (\BBB(\alpha) + q\alpha) \end{equation} for every $q\in \RR$. \item The domain of $\BBB(\alpha)$ is bounded by the following: \begin{align} \label{eqn:amin} \alpha_\mathrm{min} &= \inf \{ \alpha\in \RR \mid T_\BBB(q) \geq q\alpha \text{ for all } q \}, \\ \label{eqn:amax} \alpha_\mathrm{max} &= \sup \{ \alpha\in \RR \mid T_\BBB(q) \geq q\alpha \text{ for all } q \}, \end{align} That is, $\Kb{\alpha} = \emptyset$ for every $\alpha<\alpha_\mathrm{min}$ and every $\alpha>\alpha_\mathrm{max}$. \item Suppose that $T_\BBB$ is $\CCC^r$ on $(q_1,q_2)$ for some $r\geq 1$, and that for each $q\in(q_1,q_2)$, there exists a (not necessarily unique) equilibrium state $\nu_q$ for the potential function $q\ph$. Let $\alpha_1 = D^+T_\BBB(q_1)$ and $\alpha_2 = D^-T_\BBB(q_2)$; then \begin{equation}\label{eqn:BisTL} \BBB(\alpha)=\Tb^{L_1}(\alpha) = \inf_{q\in\RR} (T_\BBB(q) - q\alpha) \end{equation} for all $\alpha\in (\alpha_1,\alpha_2)$. In particular, $\BBB(\alpha)$ is strictly concave on $(\alpha_1,\alpha_2)$, and $\CCC^r$ except at points corresponding to intervals on which $T_\BBB$ is affine. \end{enumerate} \end{theorem} Observe that the first two statements hold for \emph{every} continuous map $f$, without any assumptions on the system, thermodynamic or otherwise. For discontinuous potentials in $\AAA_f$, these are the first rigorous multifractal results of any sort known to the author. Using the maps $A$ and $Q$ introduced above, Part III can be stated as follows: if $T_\BBB$ is $\CCC^r$ on an open interval $I_Q$ and equilibrium states exist for all $q\in I_Q$, then~\eqref{eqn:BisTL} holds for all $\alpha\in A(I_Q)$. If in addition $T_\BBB$ is strictly convex on $I_Q$, then $\BBB(\alpha)$ is $\CCC^r$ on $A(I_Q)$. We will show later that if the entropy map is upper semi-continuous, then the conclusion of Part III holds at $\alpha_1$ and $\alpha_2$ as well. We will also see (Proposition~\ref{prop:unique-works}) that existence of a \emph{unique} equilibrium state on an interval $(q_1,q_2)$ is enough to guarantee differentiability, and hence to apply Theorem~\ref{thm:birkhoff}. As shown in Example~\ref{eg:vw} below, though, we may have differentiability without uniqueness. \section{Phase transitions and generalisations of Theorem~\ref{thm:birkhoff}}\label{sec:gen} \begin{figure}[tbp] \includegraphics{no-phase-transition.eps} \caption{The Birkhoff spectrum for a map with no phase transitions.} \label{fig:no-phase-transition} \end{figure} If $T_\BBB$ is continuously differentiable for all $q$, then we obtain the complete Birkhoff spectrum, as shown in Figure~\ref{fig:no-phase-transition}. However, there are many physically interesting systems which display \emph{phase transitions}---that is, values of $q$ at which $T_\BBB$ is non-differentiable. For example, if $f\colon [0,1]\to [0,1]$ is the Manneville--Pomeau map and $\ph$ is the geometric potential $\log \abs{f'}$, then $T_\BBB$ is as shown in Figure~\ref{fig:phase-transition}~\cite{kN00}; in particular, $T_\BBB$ is not differentiable at $q_0$. Thus Theorem~\ref{thm:birkhoff} gives the Birkhoff spectrum on the interval $[\alpha_1,\alpha_2]$, where $\alpha_1 = \lim_{q\to q_0^+} T_\BBB'(q)$, but says nothing about the interval $[0,\alpha_1)$, on which $\Tb^{L_1}(\alpha) = -q_0 \alpha$ is linear. \begin{figure}[tbp] \includegraphics{phase-transition.eps} \caption{A phase transition in the Manneville--Pomeau map.} \label{fig:phase-transition} \end{figure} In fact, it is known that for this particular example, we have $\BBB(\alpha) = \Tb^{L_1}$ even on the linear stretch corresponding to the point of non-differentiability of $T_\BBB$~\cite{kN00}. However, this is not universally the case, as may be seen by ``gluing together'' two unrelated maps. Consider two maps $f_1\colon X_1 \to X_1$ and $f_2\colon X_2 \to X_2$, where $X_1$ and $X_2$ are disjoint, and suppose that the thermodynamic functions are as shown in Figure~\ref{fig:non-transitive}. Let $X = X_1 \cup X_2$, and define a map $f\colon X\to X$ such that the restriction of $f$ to $X_i$ is $f_i$ for $i=1,2$. Then $T_\BBB(q) = P^*(q\ph) = \max\{P_1^*(q\ph|_{X_1}), P_2^*(q\ph|_{X_2})\}$, where $P_i^*$ denotes the pressure of $f_i$, and furthermore $\BBB(\alpha)$ is the maximum of $h_\mathrm{top} (\Kb{\alpha}\cap X_1)$ and $h_\mathrm{top} (\Kb{\alpha}\cap X_2)$. Thus $T_\BBB$ is non-differentiable at $q=0$, which corresponds to the interval $[\alpha_2,\alpha_3]$ on which $\Tb^{L_1}$ is constant. Applying Theorem~\ref{thm:birkhoff} to each of the subsystems $f_i$, we see that $\BBB(\alpha)=\Tb^{L_1}(\alpha)$ on $[\alpha_1,\alpha_2]$ and $[\alpha_3,\alpha_4]$, but that the two are not equal on $(\alpha_2,\alpha_3)$, and that $\BBB(\alpha)$ is not concave on this interval. \begin{figure}[tbp] \includegraphics{non-transitive.eps} \caption{A different sort of phase transition.} \label{fig:non-transitive} \end{figure} \begin{example}\label{eg:vw} Given $m,n\in \NN$, let $(X_1,f_1) = (\Sigma_m^+,\sigma)$ and $(X_2,f_2)=(\Sigma_n^+,\sigma)$ be the full one-sided shifts on $m$ and $n$ symbols, respectively, and construct $f\colon X\to X$ as above, where $X=X_1 \cup X_2$. Choose two vectors $v\in \RR^m$ and $w\in \RR^n$, and let $\ph\colon X\to \RR$ be given by \[ \ph(x) = \begin{cases} v_{x_1} &x=x_1 x_2 \dots \in X_1 = \Sigma_m^+, \\ w_{x_1} &x=x_1 x_2 \dots \in X_2 = \Sigma_n^+. \end{cases} \] Then an easy computation using the classical definition of pressure and the variational principle shows that \begin{multline*} T_\BBB(q) = P^*(q\ph) = \max(P_1^*(q\ph), P_2^*(q\ph)) \\ = \max\left(\log \left(\sum_{i=1}^m e^{qv_i}\right), \log \left(\sum_{j=1}^n e^{qw_j} \right) \right). \end{multline*} In particular, we see that $P_1^*(0) = \log m$ and $P_2^*(0) = \log n$, and also that \begin{equation}\label{eqn:derivatives} \begin{aligned} \frac{d^k}{dq^k}P_1^*(q\ph)|_{q=0} &= \log\left(\sum_i v_i^k\right), \\ \frac{d^k}{dq^k}P_2^*(q\ph)|_{q=0} &= \log\left(\sum_j w_j^k\right). \end{aligned} \end{equation} By judicious choices of $v$ and $w$, we can observe a variety of behaviours in the Birkhoff spectrum $\BBB(\alpha)$. If $m=n$ but $\sum_i v_i \neq \sum_j w_j$, we obtain the picture shown in Figure~\ref{fig:non-transitive}. If $m=n$ and $\sum_i v_i = \sum_j w_j$, but $\sum_i v_i^2 > \sum_j w_j^2$, then the two pressure functions $P_1^*(q\ph)$ and $P_2^*(q\ph)$ are tangent at $q=0$, corresponding to the existence of two ergodic measures of maximal entropy (one on $X_1$ and one on $X_2$), but for values of $q$ near $0$, there is a unique equilibrium state supported on $X_1$. Finally, if $m=n$ and $\sum_i v_i^k = \sum_j w_j^k$ for $k=1,2$, but not for $k=3$, then the two pressure functions are still tangent at $q=0$, but now the equilibrium state passes from $X_1$ to $X_2$ as $q$ passes through $0$. Despite this transition and the non-uniqueness of the measure of maximal entropy, the pressure function $T_\BBB$ is still differentiable at $0$. \end{example} Having seen two very different manifestations of phase transitions (Figures~\ref{fig:phase-transition} and~\ref{fig:non-transitive}), we see that any generalisation of Theorem~\ref{thm:birkhoff} that treats phase transitions must somehow distinguish between these two sorts of behaviour. The key difference is that in the first case, the system $f\colon X\to X$ can be approximated from within by a sequence of subsystems $X_n$ on which there is no phase transition---that is, the following condition holds~\cite{kN00,GR09}: \begin{description} \item[(A)] There exists a sequence of compact $f$-invariant subsets $X_n\subset X$ such that the pressure function $q\mapsto P_{X_n}^*(q\ph)$ is continuously differentiable for all $q\in \RR$ (and equilibrium states exist), and furthermore, \begin{equation}\label{eqn:pressures-converge} \lim_{n\to\infty} P_{X_n}^*(q\ph) = P^*(q\ph). \end{equation} \end{description} This condition fails for the example in Figure~\ref{fig:non-transitive}, in which the phase transition represents a jump from one half of the system to the other half, which is disconnected from the first, rather than an escaping of measures to an adjacent fixed point. Using Condition \textbf{(A)}, we can state a general theorem which extends Theorem~\ref{thm:birkhoff} to maps for which $T_\BBB$ has points of non-differentiability. \begin{theorem}\label{thm:phase} Let $X$ be a compact metric space, $f\colon X\to X$ be continuous, and $\ph\in \AAA_f$. If Condition \textbf{(A)}\ holds, then we have~\eqref{eqn:BisTL} for all $\alpha\in (\alpha_\mathrm{min},\alpha_\mathrm{max})$. \end{theorem} As mentioned just before Theorem~\ref{thm:birkhoff}, the key property of potentials $\ph \in \AAA_f$ is that weak* convergence to an invariant measure implies convergence of the integrals of $\ph$; this is the only place in the proof where we use the requirement that $\ph$ lie in $\AAA_f$. For potentials outside of $\AAA_f$, we can try to regain approximate convergence results at certain relevant measures by using the topological entropy of $\overline{\CCC(\ph)}$ to give a bound on how much weight a neighbourhood of $\overline{\CCC(\ph)}$ carries. To this end, given $h\geq 0$, consider the set \[ I_A(h) = \{\alpha\in \RR \mid \Tb^{L_1}(\alpha) > h\}, \] and also its counterpart \[ I_Q(h) = Q(I_A(h)). \] Geometrically, $I_Q(h)$ may be described as the set of values $q\in \RR$ such that there is a line through $(q,T_\BBB(q))$ that lies on or beneath the graph of $T_\BBB$ and intersects the $y$-axis somewhere above $(0,h)$. \begin{theorem}\label{thm:high-entropy} Let $X$ be a compact metric space, $f\colon X\to X$ be continuous, and $\ph\colon X\to \RR$ be measurable and bounded (above and below). Let $\CCC(\ph)$ be the set of discontinuities of $\ph$, and let $h_0 = \underline{Ch}_\mathrm{top}(\CCC(\ph))$. Then \begin{enumerate}[I.] \item For every $q\in I_Q(h_0)$, we have the following version of~\eqref{eqn:TisBL}: \begin{equation}\label{eqn:TisBL2} T_\BBB(q) = \sup_{\alpha\in I_A(h_0)} (\BBB(\alpha) + q\alpha). \end{equation} \item $\BBB(\alpha) \leq h_0$ for every $\alpha\notin I_A(h_0)$. \item Suppose that $T_\BBB$ is $\CCC^r$ on $(q_1,q_2)\subset Q(h_0)$ for some $r\geq 1$, and that for each $q\in (q_1,q_2)$ there exists a (not necessarily unique) equilibrium state $\nu_q$ for the potential function $q\ph$. Then~\eqref{eqn:BisTL} holds for all $\alpha\in (\alpha_1,\alpha_2) = A((q_1,q_2))$. \end{enumerate} \end{theorem} Finally, although we are not yet able to give a complete treatment of unbounded potential functions, we can show that everything works if our potential function is bounded below and we only consider $q\leq 0$. \begin{theorem}\label{thm:singularity} Let $X$ be a compact metric space, $f\colon X\to X$ be continuous, and $\ph\colon X\to \RR \cup \{+\infty\}$ be continuous where finite (and hence bounded below). Let $\alpha_0 D^-T_\BBB(0)$, and let $\alpha_\mathrm{min}$ be given by~\eqref{eqn:amin}, so $(\alpha_\mathrm{min},\alpha_0) = A((-\infty,0))$. Then \begin{enumerate}[I.] \item For every $q\leq 0$,~\eqref{eqn:TisBL} holds. \item For $\alpha<\alpha_\mathrm{min}$, we have $\Kb{\alpha} = \emptyset$. \item Suppose that $T_\BBB$ is $\CCC^r$ on $(q_1,q_2)$ for some $r\geq 1$ and $q_1 < q_2\leq 0$, and that for each $q\in (q_1,q_2)$ there exists a (not necessarily unique) equilibrium state $\nu_q$ for the potential $q\ph$. Then~\eqref{eqn:BisTL} holds for all $\alpha\in (\alpha_1,\alpha_2) = A((q_1,q_2))$. \end{enumerate} \end{theorem} An analogous result holds for $q\geq 0$ if $\ph$ is bounded above but not below. Also, as with Theorem~\ref{thm:birkhoff}, Part III extends to the endpoints $\alpha_i$ if the entropy map is upper semi-continuous. \section{Conformal maps and Lyapunov spectra}\label{sec:conformal} \begin{definition} We say that a continuous map $f\colon X\to X$ is \emph{conformal} with factor $a(x)$ if for every $x\in X$ we have \begin{equation}\label{eqn:conformal} a(x) = \lim_{y\to x} \frac{d(f(x),f(y))}{d(x,y)}, \end{equation} where $a\colon X\to [0,\infty)$ is continuous. A point $x\in X$ is a \emph{critical point} of $f$ if $a(x)=0$. We denote the Birkhoff sums of $\log a$ by \[ \lambda_n(x) = \frac 1n S_n (\log a)(x), \] and consider the lower and upper limits \[ \underline{\lambda}(x) = \llim_{n\to\infty} \lambda_n(x), \qquad \overline{\lambda}(x) = \ulim_{n\to\infty} \lambda_n(x). \] If they agree (that is, if the limit exists), we write \[ \lambda(x) = \lim_{n\to\infty} \lambda_n(x) \] for the \emph{Lyapunov exponent} at $x$. Given a measure $\mu\in\MMM(X)$ we define the Lyapunov exponent of $\mu$ as \[ \lambda(\mu) = \int_X \lambda(x) \,d\mu(x). \] If $\mu$ is ergodic, then $\lambda(\mu)=\lambda(x)$ for $\mu$-almost every $x\in X$. \end{definition} Note that in the case where $X$ is a smooth Riemannian manifold, the definition of conformality may be restated as the requirement that $Df(x)$ is $a(x)$ times some isometry, and the definition of Lyapunov exponent becomes the usual one from smooth ergodic theory. In particular, if $X$ is one-dimensional, then any differentiable map is conformal. Denote by $\mathbf{B}$ the set of all points in $X$ which satisfy (at least) one of the following two conditions. \begin{description} \item[(B1)] \emph{Bounded contraction:} $\inf \{ S_n(\log a)(f^k(x)) \mid k,n\in\NN \} > -\infty$. Note that if $a(x) \geq 1$ for all $x\in X$, then $f$ has no contraction whatsoever (although the expansion may not be uniform), and so every point has bounded contraction. \item[(B2)] \emph{Lyapunov exponent exists:} $\underline{\lambda}(x)=\overline{\lambda}(x)$. \end{description} The following lemma is proved in~\cite{vC09a}, and shows that we can dynamically generate metric balls using conformal maps without critical points. We will need this later for the results on $\DDD(\alpha)$ in Section~\ref{sec:wkgibbs}. \begin{lemma}\label{lem:well-behaved} Let $X$ be a compact metric space and $f\colon X\to X$ be continuous and conformal with factor $a(x)$. Suppose that $f$ has no critical points; that is, that $a(x)>0$ for all $x\in X$. Then given any $x\in \mathbf{B}$ and $\eps>0$, there exists $\delta=\delta(\eps)>0$ and $\eta=\eta(x)>0$ such that for every $n$, \begin{equation}\label{eqn:diamball} B\left(x,\eta\delta e^{-n(\lambda_n(x) + \eps)}\right) \subset B(x,n,\delta) \subset B\left(x,\delta e^{-n(\lambda_n(x) -\eps)}\right). \end{equation} \end{lemma} Using this result, it is shown in~\cite{vC09a} that if $f$ is a conformal map without critical points, then given $Z\subset X$ and $\alpha>0$ such that \begin{equation}\label{eqn:lyapconst} \underline{\lambda}(x) = \overline{\lambda}(x) = \alpha \end{equation} for every $x\in Z$, the Hausdorff dimension and topological entropy of $Z$ are related by \begin{equation}\label{eqn:htopdimh} \dim_H Z = \frac 1\alpha h_\mathrm{top} Z. \end{equation} Recall that the level sets $\Kb{\alpha}$ for the Birkhoff averages of the geometric potential $\ph = \log a$ are precisely the level sets $\Kl{\alpha}$ for the Lyapunov exponents of $f$, and thus $\LLL_E(\alpha) = \BBB(\alpha)$ is determined by $T_\BBB$ using Theorem~\ref{thm:birkhoff}. Since every point $x\in \Kl{\alpha}$ satisfies~\eqref{eqn:lyapconst}, we may apply~\eqref{eqn:htopdimh} and obtain \begin{equation}\label{eqn:dimlyap} \LLL_D(\alpha) = \frac 1\alpha \LLL_E(\alpha) \end{equation} for all $\alpha>0$. Thus both Lyapunov spectra can be determined in terms of the Legendre transform of $T_\BBB$, provided equilibrium states exist and $T_\BBB$ is differentiable. We stress that since $\LLL_D(\alpha)$ is not given by a Legendre transform, but is obtained by a rescaling, it may not be convex---see~\cite{IK09} for examples where this occurs. \section{Entropy and dimension spectra of weak Gibbs measures}\label{sec:wkgibbs} The two remaining multifractal spectra with which we are concerned---the entropy spectrum and the dimension spectrum---are both defined in terms of a measure $\mu$. In order to relate these spectra to the thermodynamic quantities associated with a potential $\ph$, we need a relationship between the local properties of $\mu$ and the Birkhoff averages of $\ph$. This is provided by the notion of a weak Gibbs measure. \begin{definition} Given a compact metric space $X$, a continuous map $f\colon X\to X$, and a potential $\ph\colon X\to \RR$ (not necessarily continuous), we say that a Borel probability measure $\mu$ is a \emph{weak Gibbs measure} for $\ph$ with constant $P\in \RR$ if for every $x\in X$ and $\delta>0$ there exists a sequence $M_n = M_n(x,\delta) > 0$ such that \begin{equation}\label{eqn:Gibbs} \frac 1{M_n} \leq \frac{\mu(B(x,n,\delta))}{\exp(-nP + S_n\ph(x))} \leq M_n \end{equation} for every $n\in \NN$, where we require the following growth condition on $M_n$ to hold for every $x\in X$: \begin{equation}\label{eqn:tempered} \lim_{\delta\to 0} \ulim_{n\to\infty} \frac 1n \log M_n(x,\delta) = 0. \end{equation} \end{definition} There are various definitions in the literature of Gibbs measures of one sort or another; most of these definitions agree in spirit, but differ in some slight details. We note the differences between the above definition and other definitions in use. \begin{enumerate}[(1)] \item The classical definition (see~\cite{rB75}) requires $M_n$ to be bounded, not just to have slow growth, as we require here. In that case the sequence $M_n$ can be (and is) replaced by a single constant $M$. The notion of a weak Gibbs measure, for which the constant can vary slowly in $n$, is used in~\cite{mY00,mK01,FO03,JR09}, among others. \item The above definitions all require the constant $M$ to be independent of $x$, whereas we require no such uniformity. Furthermore, they are given in terms of cylinder sets rather than Bowen balls; we follow~\cite{VV08} in using the latter, as this is what we need for the multifractal analysis. \item Certain authors only require~\eqref{eqn:Gibbs} to hold for $\mu$-a.e.\ $x\in X$~\cite{mY00,VV08}. In order to do the multifractal analysis, we need conditions which hold everywhere, not just almost everywhere, and so we require~\eqref{eqn:Gibbs} for \emph{every} point $x\in X$. \item Following Kesseb\"ohmer~\cite{mK01}, we do not \emph{a priori} require that a weak Gibbs measure be $f$-invariant. Weak Gibbs measures exist for \emph{any} continuous function $\ph$ on a one-sided shift space~\cite{mK01}, but it is not the case that such measures can always be taken to be invariant. \end{enumerate} We have given the definition in the above form because~\eqref{eqn:Gibbs} is reminiscent of the usual definition of Gibbs measure. For our purposes, an alternate form of~\eqref{eqn:Gibbs} will be more useful: \begin{equation}\label{eqn:Gibbs2} \left\lvert -\frac 1n \log \mu(B(x,n,\delta)) + \frac 1n S_n \ph(x) - P \right\rvert \leq \frac 1n\log M_n(x,\delta) \to 0, \end{equation} where the limit is taken as $n\to \infty$ and then as $\delta \to 0$. Given an invariant weak Gibbs measure, it follows from~\eqref{eqn:Gibbs2} that $h_\mu(x)$ exists if and only if $\ph^+(x)$ exists, and that in this case \begin{equation}\label{eqn:handph} h_\mu(x) + \ph^+(x) = P. \end{equation} If $\ph$ is continuous, then dimensional arguments from~\cite{yP98} show that $P$ is equal to the topological pressure $P_X(\ph)$, and thus the variational principle shows that it is equal to $P^*(\ph)$. Integrating~\eqref{eqn:handph} with respect to $\mu$, we obtain $P^*(\ph) = h(\mu) + \int \ph \,d\mu$, hence $\mu$ is an equilibrium state. Thus a weak Gibbs measure is an equilibrium state, just as in the classical case. For any equilibrium state, the Brin--Katok entropy formula and the Birkhoff ergodic theorem together imply that~\eqref{eqn:handph} holds almost everywhere with $P=P^*(\ph)$; our definition of weak Gibbs measure boils down to requiring that it hold \emph{everywhere}, without placing any extra requirements on uniformity or rate of convergence. Writing $\ph_1(x) = \ph(x) - P^*(\ph)$, we observe that \begin{equation}\label{eqn:levelsets} \Kb{\alpha}(\ph_1) = K_{-\alpha}^\mathcal{E}, \end{equation} and we may thus obtain $\EEE(\alpha)$ as a Legendre transform of the following function: \[ T_\EEE(q) = P^*(q\ph_1). \] Once again, convexity of $T_\EEE$ is immediate from the definition of $P^*$. The following theorem is a direct consequence of Theorem~\ref{thm:birkhoff} and~\eqref{eqn:handph}; because of the change of sign in~\eqref{eqn:levelsets}, we must use the following versions of the Legendre transform: \begin{equation}\label{eqn:TLSL} \begin{aligned} T^{L_3}(\alpha) &= \inf_{q\in\RR} (T(q) + q\alpha), \\ S^{L_4}(q) &= \sup_{\alpha\in\RR} (S(\alpha) - q\alpha). \end{aligned} \end{equation} Note that there is a corresponding change of sign in the definitions of the maps $A$ and $Q$. \begin{theorem}[The entropy spectrum for local entropies]\label{thm:entropy} Let $X$ be a compact metric space, $f\colon X\to X$ be continuous, and $\ph\in \AAA_f$. Then if $\mu$ is a weak Gibbs measure for $\ph$, we have the following: \begin{enumerate}[I.] \item $T_\EEE$ is the Legendre transform of the entropy spectrum: \begin{equation}\label{eqn:TisEL} T_\EEE(q) = \BBB^{L_4}(q) = \sup_{\alpha\in\RR} (\EEE(\alpha) - q\alpha) \end{equation} for every $q\in \RR$. \item The domain of $E$ is bounded by the following: \begin{align*} \alpha_\mathrm{min} &= \inf \{ \alpha\in \RR \mid T_\EEE(q) \geq -q\alpha \text{ for all } q \}, \\ \alpha_\mathrm{max} &= \sup \{ \alpha\in \RR \mid T_\EEE(q) \geq -q\alpha \text{ for all } q \}, \end{align*} That is, $\Ke{\alpha} = \emptyset$ for every $\alpha<\alpha_\mathrm{min}$ and every $\alpha>\alpha_\mathrm{max}$. \item Suppose that $T_\EEE$ is $\CCC^r$ on $(q_1,q_2)$ for some $r\geq 1$, and that for each $q\in(q_1,q_2)$, there exists a (not necessarily unique) equilibrium state $\nu_q$ for the potential function $q\ph_1$. Let $\alpha_1 = -D^+T_\EEE(q_1)$ and $\alpha_2 = D^-T_\EEE(q_2)$. Then \begin{equation}\label{eqn:EisTL} \EEE(\alpha)=\Te^{L_3}(\alpha) = \inf_{q\in\RR} (T_\EEE(q) + q\alpha) \end{equation} for all $\alpha\in (\alpha_2,\alpha_1)$; in particular, $E$ is strictly concave on $(\alpha_2,\alpha_1)$, and $\CCC^r$ except at points corresponding to intervals on which $T_\EEE$ is affine. \end{enumerate} \end{theorem} In the case where $f$ is conformal, we prove the analogous result for the dimension spectrum. We will need to eliminate points at which the Birkhoff averages of $\log a$ cluster around zero along a sequence of times at which the local entropy of $\mu$ is also negligible; that is, the following set: \begin{equation}\label{eqn:Z} \mathbf{Z}(\mu) = \left\{ x\in X \,\Big|\, \lim_{\delta\to 0} \llim_{n\to\infty} \left\lvert \frac 1n \log\mu(B(x,n,\delta)) \right\rvert + \left\lvert \frac 1n S_n \log a(x) \right\rvert = 0 \right\}. \end{equation} When $\mu$ is a weak Gibbs measure for $\ph$, we have \begin{equation}\label{eqn:Z2} \mathbf{Z}(\mu) = \left\{ x\in X \,\Big|\, \llim_{n\to\infty} \left\lvert \frac 1n S_n \ph_1(x) \right\rvert + \left\lvert \frac 1n S_n \log a(x) \right\rvert = 0 \right\}. \end{equation} In the context of Theorem~\ref{thm:dimension}, we will suppress the dependence on $\mu$ and simply write $\mathbf{Z} = \mathbf{Z}(\mu)$. We will see that the set $\mathbf{Z}$ contains all points $x$ for which $\underline{\lambda}(x) = 0$ but $\ud_\mu(x)<\infty$; these are the only points our methods cannot deal with. In many cases we do not lose much by neglecting them; for example, if $\sup \ph - \inf \ph < h(\mu)$, then \[ \ulim_{n\to\infty} \frac 1n S_n \ph_1(x) < 0 \] for every $x\in X$, and so $\mathbf{Z} = \emptyset$. Even in cases when $\mathbf{Z}$ is non-empty, it often has zero Hausdorff dimension~\cite{JR09}. The remaining set of ``good'' points will be denoted by \begin{equation}\label{eqn:X'} X' = X \setminus \mathbf{Z}. \end{equation} In the definition of $\DDD(\alpha)$, we adopt the convention that $\DDD(\alpha) = -\infty$ if $\Kd{\alpha} \subset \mathbf{Z}$. Since there may be points at which $\mu$ has infinite pointwise dimension, we also include the value $\alpha=+\infty$ in~\eqref{eqn:TLSL}, and follow the convention that if $K_\infty^\mathcal{D} \cap X' \neq \emptyset$, then $\DDD^{L_4}(q) = +\infty$ for all $q<0$. Now consider the centred potential $\ph_1(x) = \ph(x) - P^*(\ph)$. Define a family of potentials by \begin{equation}\label{eqn:conformal2} \ph_{q,t}(x) = q\ph_1(x) - t\log a(x). \end{equation} We will be particularly interested in the potentials with zero pressure; we would like to define a function $T_\DDD(q)$ by the equation \begin{equation}\label{eqn:Pzero} P^*\left(\ph_{q,T_\DDD(q)}\right) = 0. \end{equation} Formally, we write \begin{equation}\label{eqn:Td} T_\DDD(q) = \inf \{ t\in\RR \mid P^*(\ph_{q,t}) \leq 0 \} = \sup\{t\in \RR \mid P^*(\ph_{q,t}) > 0\}; \end{equation} by continuity of $P^*$, $T_\DDD(q)$ solves~\eqref{eqn:Pzero} if it is finite, but is not necessarily the unique solution of~\eqref{eqn:Pzero}. (Indeed, there may be values of $q$ for which $P^*(\ph_{q,t}) = 0$ for all $t>T_\DDD(q)$.) For $T_\DDD(q)<\infty$ we write $\ph_q = \ph_{q,T_\DDD(q)}$, and observe that~\eqref{eqn:Pzero} may be written as $P^*(\ph_q)=0$. Given $\eta>0$ and $I_Q\subset \RR$, we will need to consider the following region lying just under the graph of $T_\DDD(q)$: \[ R_\eta(I_Q) = \{(q,t)\in \RR^2 \mid q\in I_Q, T_\DDD(q) - \eta < t < T_\DDD(q) \}. \] We can now state a general result regarding the dimension spectrum. \begin{theorem}[The dimension spectrum for pointwise dimensions]\label{thm:dimension} Let $X$ be a compact metric space with $\dim_H X < \infty$, and let $f\colon X\to X$ be continuous and conformal with continuous non-vanishing factor $a(x)$. Suppose that $\mathbf{B}=X$ and that $\lambda(\nu)\geq 0$ for every $\nu\in \MMM^f(X)$. Let $\mu\in \MMM^f(X)$ be a weak Gibbs measure for a continuous potential $\ph$. Finally, suppose that $\dim_H \mathbf{Z} = 0$. Then we have the following. \begin{enumerate}[I.] \item $T_\DDD$ is the Legendre transform of the dimension spectrum: \begin{equation}\label{eqn:TisDL} T_\DDD(q) = \DDD^{L_4}(q) = \sup_{\alpha\in\RR} (\DDD(\alpha) - q\alpha) \end{equation} for every $q\in \RR$. \item Neglecting points in $\mathbf{Z}$, the domain of $\mathcal{D}$ is bounded by the following: \begin{align*} \alpha_\mathrm{min} &= \inf \{ \alpha\in \RR \mid T_\DDD(q) \geq -q\alpha \text{ for all } q \}, \\ \alpha_\mathrm{max} &= \sup \{ \alpha\in \RR \mid T_\DDD(q) \geq -q\alpha \text{ for all } q \}, \end{align*} That is, $\Kd{\alpha} \cap X' = \emptyset$ for every $\alpha<\alpha_\mathrm{min}$ and every $\alpha>\alpha_\mathrm{max}$. \item Suppose $I_Q = (q_1,q_2)$ and $\eta>0$ are such that for every $(q,t)\in R_\eta(I_Q)$, the potential $\ph_{q,t}$ has a (not necessarily unique) equilibrium state, and that the map $(q,t) \mapsto P^*(\ph_{q,t})$ is $\CCC^r$ on $R_\eta(I_Q)$ for some $r\geq 1$. Then we have \begin{equation}\label{eqn:DisTL} \DDD(\alpha) = \Td^{L_3}(\alpha) = \inf_{q\in\RR} (T_\DDD(q) + q\alpha) \end{equation} for all $\alpha\in (\alpha_2,\alpha_1) = A(I_Q)$; in particular, $\mathcal{D}$ is strictly concave on $(\alpha_2,\alpha_1)$, and $\CCC^r$ except at points corresponding to intervals on which $T_\DDD$ is affine. \end{enumerate} \end{theorem} We will see in the proof that the requirement on existence of equilibrium states for $\ph_{q,t}$ with $(q,t)\in R_\eta(I_Q)$ can be replaced by the condition that there exist equilibrium states $\nu_q$ for $\ph_q = \ph_{q,T_\DDD(q)}$ such that $\lambda(\nu_q) > 0$. However, such measures do not necessarily exist, while upper semi-continuity of the entropy is enough to guarantee the existence of the measures required in the theorem. If we do have equilibrium states $\nu_q$ with $\lambda(\nu_q) > 0$, then in Part III of the theorem, the requirement that $(q,t)\mapsto P^*(\ph_{q,t})$ be $\CCC^r$ on $R_\eta(I_Q)$ can be replaced by the condition that $T_\DDD$ be $\CCC^r$ on $I_Q$. \section{Remarks}\label{sec:rmk} We first discuss conditions under which the hypotheses of Theorem~\ref{thm:birkhoff} and the results in Section~\ref{sec:gen} are satisfied, before turning our attention to weak Gibbs measures and Theorem~\ref{thm:entropy}, and finally the more delicate case of Theorem~\ref{thm:dimension}. Throughout this section, $T$ will refer to any or all of $T_\BBB$, $T_\EEE$, and $T_\DDD$, as needed. We make general remarks in this section, deferring specific examples and applications until Section~\ref{sec:app}. \subsection{Birkhoff spectrum---continuous potentials} Parts I and II of Theorem~\ref{thm:birkhoff} and~\ref{thm:entropy} do not place any thermodynamic requirements on the function $T=T_\BBB$, and thus hold in full generality. There are two thermodynamic requirements in Part III---existence of an equilibrium state, and differentiability of $T$. The latter is used in order to guarantee the existence of values $q\in \RR$ for which $T_\BBB'(q)$ exists, and hence $A(q) = \{T_\BBB'(q)\}$ is a singleton. In fact, because $T$ is continuous and convex, $A(q)$ is a singleton for all but at most countably many values of $q$, and consequently, once existence of equilibrium states is established, it follows that the Birkhoff spectrum is equal to the Legendre transform of the pressure function everywhere except possibly on some countable union of intervals, on each of which that Legendre transform is affine and gives an upper bound for $\BBB(\alpha)$. Existence of equilibrium states is easy to verify in the following rather common setting. \begin{definition} The entropy map $\mu\mapsto h(\mu)$ is \emph{upper semi-continuous} if for every sequence $\mu_n\in\MMM(X)$ which converges to $\mu$ in the weak* topology, we have \[ \ulim_{n\to\infty} h(\mu_n) \leq h(\mu). \] \end{definition} If the entropy map is upper semi-continuous and $\ph$ is continuous, then the map \[ \mu \mapsto h(\mu) + \int q\ph \,d\mu \] is upper semi-continuous for every $q\in\RR$, and thus attains its maximum. In particular, there exists an equilibrium state for every $q\ph$. \begin{definition} $f$ is \emph{expansive} if there exists $\eps>0$ such that for all $x\neq y$ there exists $n\in \ZZ$ (if $f$ is invertible) or $n\in \NN$ (if $f$ is non-invertible) such that $d(f^n(x),f^n(y))\geq \eps$. \end{definition} For expansive homeomorphisms, the entropy map $\mu\mapsto h_\mu(f)$ is upper semi-continuous~\cite[Theorem 8.2]{pW75},\foot{What happens if $f$ is non-invertible?} and so existence is guaranteed for continuous $\ph$. Similarly, the entropy map is upper semi-continuous for $\CCC^\infty$ maps of compact smooth manifolds~\cite{sN89}, and we once again get existence for free. \begin{proposition}\label{prop:unique-works} Let $X$ be a compact metric space, $f\colon X\to X$ a continuous map, and $\ph\in \AAA_f$. Suppose that the entropy map is upper semi-continuous and that there exists an interval $(q_1,q_2)\subset \RR$ such that for every $q\in (q_1,q_2)$, the potential $q\ph$ has a unique equilibrium state. Then $T_\BBB$ is $\CCC^1$ on $(q_1,q_2)$. \end{proposition} \begin{proof} Suppose for a contradiction that the pressure function $q\mapsto P^*(q\ph)$ is not differentiable at $q_0\in (q_1,q_2)$. Let $\mu_n^-$ be the unique equilibrium state for $(q-\frac 1n)\ph$, and let $\mu^-$ be a weak* limit of some subsequence $\mu_{n_j}^-$. By upper semi-continuity and Proposition~\ref{prop:convergence} below, we have \begin{multline*} h(\mu^-) + \int q\ph\,d\mu^- \geq \ulim_{n_j\to\infty} h(\mu_{n_j})+ \int q\ph\,d\mu_{n_j}^- \\ = \ulim_{n_j\to\infty} P^*\left(\left(q-\frac 1{n_j}\right)\ph\right) = P^*(q\ph). \end{multline*} Thus $\mu^-$ is an equilibrium state for $q\ph$ with \[ \int q\ph\,d\mu^- = D^-T_\BBB(q) = \lim_{q'\to q^-} T_\BBB'(q') \] by Proposition~\ref{prop:ruelle} below. Similarly, one can construct an equilibrium state $\mu^+$ such that $\int q\ph\,d\mu^+$ is the right derivative of $T_\BBB$ at $q$. If the two derivatives do not agree, then we have two distinct equilibrium states for $q\ph$, a contradiction. \end{proof} Using Proposition~\ref{prop:unique-works}, one approach to verifying the hypotheses of Theorem~\ref{thm:birkhoff} for a map with upper semi-continuous entropy is to show that the equilibrium state for each $q\ph$ is unique. We also observe that in the context of Part III of Theorem~\ref{thm:birkhoff}, the construction in the proof above gives equilibrium states for $q_1\ph$ and $q_2\ph$ that are supported on the sets $K_{\alpha_1}^\BBB$ and $K_{\alpha_2}^\BBB$, respectively, and which establish~\eqref{eqn:BisTL} for the endpoints $\alpha_1$ and $\alpha_2$, just as in the proof of Proposition~\ref{prop:concave} below. \subsection{Birkhoff spectrum---discontinuous potentials} If $\ph$ is discontinuous, the map from $\MMM(X)$ to $\RR$ defined by \begin{equation}\label{eqn:intph} \mu\mapsto \int \ph\,d\mu \end{equation} is not continuous on all of $\MMM(X)$. For discontinuous potentials lying in $\AAA_f$, continuity still holds at measures in $\MMM^f(X)$ by Proposition~\ref{prop:convergence} below, which suffices for all the proofs here. However, if $\ph\notin \AAA_f$, then there may be invariant measures at which the map is discontinuous. In particular, if $\mu(\CCC(\ph)) > 0$, then the map in~\eqref{eqn:intph} is discontinuous at $\mu$. If $\ph$ is unbounded, then it is relatively straightforward to show that the map is not continuous at \emph{any} measure in $\MMM(X)$. In many cases, it is not even enough to restrict our attention to invariant measures~\cite[Proposition 2.8]{BK98}. Thus for $\ph\notin \AAA_f$, upper semi-continuity of the entropy is not enough to guarantee existence of equilibrium states without further information. For potentials which are bounded above but not below, we observe in Proposition~\ref{prop:singularity} that the map in~\eqref{eqn:intph} is upper semi-continuous, and thus the free energy function $\mu\mapsto h(\mu) + \int \ph\,d\mu$ is upper semi-continuous as well. It follows that it attains its maximum, and we once again are guaranteed existence. This is also enough to prove Proposition~\ref{prop:unique-works} for these potentials, showing that existence and uniqueness imply differentiability of the pressure function (for the appropriate sign of $q$) if the entropy map is upper semi-continuous. \subsection{Entropy spectrum---weak Gibbs measures} There are many cases in which equilibrium states are known to have the weak Gibbs property~\eqref{eqn:Gibbs} or one which implies it. For example, equilibrium states for H\"older continuous potentials on uniformly hyperbolic systems are known to be Gibbs, as are equilibrium states for potentials satisfying a certain regularity property on expansive maps with specification~\cite{TV99}. Finally, Kesseb\"ohmer proves the existence of weak Gibbs measures for continuous potentials on symbolic space~\cite{mK01} (these measures are studied by Jordan and Rams~\cite{JR09} on parabolic interval maps). Given a weak Gibbs measure, all the above remarks regarding the Birkhoff spectrum apply to the entropy spectrum. \subsection{Dimension spectrum} Because of the geometric implications of any result regarding the dimension spectrum, we must deal with a more restricted class of systems. In particular, the present approach is completely dependent upon conformality of the map $f$; without conformality, we have no analogue of Lemma~\ref{lem:well-behaved} or Proposition~\ref{prop:localdim}. If analogues of these can be found in the non-conformal case, then it may be possible to establish a non-conformal version of the present result; however, this appears to require the use of a non-additive version of the thermodynamic formalism~\cite{lB96,FH10}. We also presently lack the tools to deal with maps with critical points. To establish an analogue of Lemma~\ref{lem:well-behaved} for such maps would require an estimate on the rate of recurrence of fairly arbitrary orbits to the critical point in order to control the distortion. The other hypotheses in Theorem~\ref{thm:dimension} are less restrictive, and are satisfied for quite general classes of maps. We discuss them briefly. $\BBB = X$. If $a(x)\geq 1$ for all $x$, then this is automatically satisfied; we do not need $a(x)>1$, nor any uniformity, and so the class of systems with this property includes Manneville--Pomeau maps and parabolic rational maps. Due to recurrence of the critical point, bounded contraction \emph{per se} cannot be expected to hold for maps with critical points; however, the requirement of bounded contraction can in fact be weakened slightly to include cases where the absolute value of the quantity in \textbf{(B1)} is not bounded, but grows sublinearly in $n+k$, which corresponds to a certain sort of slow recurrence. This approach, however, has yet to bear fruit. $\dim_H \mathbf{Z} = 0$. Points at which $\lambda(x) = 0$ and $d_\mu(x)<\infty$ are problematic for various reasons, and so we want to avoid having to deal with them. Since all such points lie in the set $\mathbf{Z}$, we can do this by neglecting $\mathbf{Z}$ in all our computations, and it turns out that this is not a very heavy price to pay. Of course if $f$ is uniformly expanding, this set is empty, but even in the non-uniformly expanding case, it is shown in~\cite{JR09} that $\mathbf{Z}$ has zero Hausdorff dimension for a class of parabolic interval maps. We also observe that if the entropy map is upper semi-continuous, then existence of equilibrium states for $\ph_{q,t}$ is guaranteed for all $q,t\in \RR$, and that uniqueness is again enough to establish differentiability of the map $(q,t)\mapsto P^*(\ph_{q,t})$, and hence to apply Theorem~\ref{thm:dimension}. \section{Applications}\label{sec:app} Before proceeding to the proofs of the theorems, we give several concrete applications. \subsection{Birkhoff spectrum} The first two parts of Theorem~\ref{thm:birkhoff} do not require \emph{any} hypotheses on the map $f$ beyond continuity, and so for every continuous map $f$ and every potential $\ph\in \AAA_f$, the pressure function $T_\BBB$ is the Legendre transform of $\BBB(\alpha)$ (and hence $\Tb^{L_1}$ is the concave hull of $\BBB(\alpha)$), and the domain of the Birkhoff spectrum is the interval $[\alpha_\mathrm{min},\alpha_\mathrm{max}]$. Similar but weaker statements hold for arbitrary bounded measurable potentials $\ph$, using Theorem~\ref{thm:high-entropy}, and for potentials with singularities using Theorem~\ref{thm:singularity}. To apply the full strength of these three theorems beyond the general remarks made so far, we need some thermodynamic information about the system. \subsubsection{Uniform hyperbolicity} In~\cite{rB75}, Bowen showed that if $M$ is a $\CCC^\infty$ Riemannian manifold and $f\colon M\to M$ is an Axiom A diffeomorphism, then any H\"older continuous potential function $\ph\colon M\to\RR$ has a unique equilibrium state. Since such maps are expansive on the hyperbolic set~\cite[Corollary 6.4.10]{KH95}, this suffices to check the hypotheses of Theorem~\ref{thm:birkhoff}, as shown in the previous section, and hence the Birkhoff spectrum is equal to the Legendre transform of the pressure function: in particular, it is concave and $\CCC^1$ (see Figure~\ref{fig:no-phase-transition}). Versions of this result may be extracted from the results in~\cite{TV99,PW01}, but Theorem~\ref{thm:birkhoff} provides a more direct proof. Non-H\"older potentials were studied by Pesin and Zhang in~\cite{PZ06} (see also~\cite{hH08}). They consider a uniformly piecewise expanding full-branched Markov map $f$ of the unit interval, and use inducing schemes and tools from the theory of countable Markov shifts to study the existence and uniqueness of equilibrium states for a large class of potentials. In particular, they give the following example of a non-H\"older potential: \begin{equation}\label{eqn:nonHolder} \ph(x) = \begin{cases} -(1-\log x)^{-\alpha} & x\in (0,1], \\ 0 & x=0. \end{cases} \end{equation} It is shown in~\cite{PZ06} that for any $\alpha>1$ and $q\in \RR$, the potential $q\ph$ has a unique equilibrium state. Since $f$ is expansive, by the comments in the previous section this suffices to check the hypotheses of Theorem~\ref{thm:birkhoff}, and we have the following result. \begin{proposition}\label{prop:nonHolder} Let $f$ be a uniformly piecewise expanding full-branched Markov map of the unit interval, and let $\ph$ be the potential function given in~\eqref{eqn:nonHolder}, $\alpha>1$. Then the Birkhoff spectrum $\BBB(\alpha)$ is smooth and concave, has domain $[\alpha_\mathrm{min},\alpha_\mathrm{max}]$, and is the Legendre transform of $T_\BBB$. \end{proposition} Indeed, Proposition~\ref{prop:nonHolder} also holds for any potential $\ph$ such that all $q\ph$ are in the class considered by Pesin and Zhang. For $0<\alpha\leq 1$, it is shown in~\cite{PZ06} that $T_\BBB$ has a phase transition at some value $q_0>0$. Applying Theorem~\ref{thm:birkhoff}, we obtain a result for the non-linear part of the Birkhoff spectrum (see Figure~\ref{fig:phase-transition}); to obtain a complete result, we would need to apply Theorem~\ref{thm:phase} by establishing Condition \textbf{(A)}. Although this remains open, one might attempt to do this by using the fact that for a potential with summable variations, the Gurevich pressure on a topologically mixing countable Markov shift $X$ is the supremum of the classical topological pressure over topologically mixing finite Markov subshifts of $X$~\cite{oS99}; these finite subshifts give natural candidates for the compact invariant sets $X_n$ in Condition \textbf{(A)}. \begin{remark} In~\cite{PS07}, Pfister and Sullivan prove a variational principle for the topological entropy of saturated sets, which include in particular the level sets $\Kb{\alpha}$, under the assumption that the system in question satisfies two properties, which they call the \emph{g-almost product property} and the \emph{uniform separation property}. Expansive systems satisfy the latter, and uniformly hyperbolic systems satisfy the former. For such systems, they prove (among other things) the following multifractal result for any continuous $\ph$~\cite[Proposition 7.1]{PS07}: \begin{equation}\label{eqn:PfSu} \BBB(\alpha) = h_\mathrm{top}(\Kb{\alpha}) = \sup\left\{ h(\mu) \,\Big|\, \mu\in \MMM^f(X), \int \ph\,d\mu = \alpha\right\}. \end{equation} Given~\eqref{eqn:PfSu}, it is not difficult to show that~\eqref{eqn:BisTL} holds, which establishes the multifractal formalism for systems with the g-almost product property and uniform separation, provided the potential is continuous. In particular, this includes the example given above, as well as some (but by no means all) of the examples mentioned below. \end{remark} \subsubsection{Parabolic maps} An important class of non-uniformly expanding maps is the Manneville--Pomeau maps, which are non-uniformly expanding interval maps with an indifferent fixed point. The primary potential of interest in this case is the geometric potential $\log\abs{f'}$, which corresponds to studying a non-H\"older potential on a \emph{uniformly} expanding interval map via an appropriate change of coordinates; thus this is closely related to the previous example. The thermodynamic properties and Lyapunov spectra of these maps were studied in~\cite{kN00,GR09}; once again, Theorem~\ref{thm:birkhoff} provides a direct proof of the multifractal results using the thermodynamic results, although as above, one would need to establish Condition \textbf{(A)}\ to deal with the linear parts of the spectrum using Theorem~\ref{thm:phase}. We also remark that a significant achievement of~\cite{GR09} is to deal with the endpoints of the spectrum ($\lambda=0$ and $\lambda=\infty$), which cannot be dealt with using the present results. Moving to two (real) dimensions, let $f\colon \overline{\CC} \to \overline{\CC}$ be a parabolic rational map of the Riemann sphere; that is, a rational map such that the Julia set $J(f)$ contains at least one indifferent fixed point (that is, a fixed point $z_0$ for which $|f'(z_0)|=1$), but does not contain any critical points. Following Makarov and Smirnov~\cite{MS00}, we say that $f$ is \emph{exceptional} if there is a finite, non-empty set $\Sigma \subset \overline{\CC}$ such that $f^{-1}(\Sigma) \setminus \Crit f = \Sigma$, where $\Crit f$ is the set of critical points of $f$. Let $\ph(z) = \log |f'(z)|$ be the geometric potential; combining the results in~\cite{MS00} with~\cite[Corollary D.1 and Theorem G]{hH08}, we see that if $f$ is non-exceptional, then the graph of the function $T_\BBB$ is as shown in Figure~\ref{fig:phase-transition}. In particular, $T_\BBB$ is analytic and strictly convex on $(q_0,\infty)$, where $q_0 = -\dim_H J(f)$, and so writing \[ \alpha_1 = D^+T_\BBB(q_0), \qquad \alpha_2 = \lim_{q \to \infty} T_\BBB'(q), \] it follows from Theorem~\ref{thm:birkhoff} that $\BBB(\alpha) = \Tb^{L_1}$ on $(\alpha_1,\alpha_2)$. Since we are dealing with the geometric potential, this is also the entropy spectrum for Lyapunov exponents, and we may apply~\eqref{eqn:dimlyap} to obtain the dimension spectrum for Lyapunov exponents, $\LLL_D(\alpha) = \frac 1\alpha \Tb^{L_1}$. This result is obtained by other methods in~\cite{GPR09}, where it is also shown that the spectra are linear on $[0,\alpha_1]$ (the dotted line in Figure~\ref{fig:phase-transition}). As before, giving an alternate proof of this using Theorem~\ref{thm:phase} would require establishing Condition \textbf{(A)}. Once again, Pfister and Sullivan's results establish the formalism for the Birkhoff spectrum here, but \emph{not} for the dimension spectrum for Lyapunov exponents, as they only consider topological entropy. \subsubsection{Other non-uniformly hyperbolic systems} The existence and uniqueness of equilibrium states for a broad class of non-uniformly expanding maps in higher dimensions was studied by Oliveira and Viana~\cite{OV08} and by Varandas and Viana~\cite{VV08}. To the best of the author's knowledge, the multifractal properties of these systems have not been studied at all, and so they provide an ideal application of Theorem~\ref{thm:birkhoff}. It does not appear to be known whether or not these systems, which may have contracting regions, satisfy specification or any other property that would imply Pfister and Sullivan's g-almost product property, and so the results of~\cite{PS07} cannot be applied. We describe the systems studied in~\cite{VV08} and use the results of that paper to apply Theorem~\ref{thm:birkhoff}. Let $M$ be a compact manifold of dimension $m$ with distance function $d$ (more generally, Varandas and Viana consider metric spaces in which the Besicovitch covering lemma holds). Let $f\colon M\to M$ be a local homeomorphism, and let $L(x)$ be a bounded function such that for every $x\in M$ there exists a neighbourhood $U_x\ni x$ such that $f_x = f|_{U_x} \colon U_x \to f(U_x)$ is invertible, with \[ d(f(y),f(z)) \geq \frac 1{L(x)} d(y,z) \] for all $y,z\in U_x$. Thus if $L(x) <1$, then $f$ is expanding at $x$, while if $L(x)\geq 1$, then $L$ controls how much contraction can happen near $x$. Assuming every point has finitely many preimages, we write $\deg_x(f) = \# f^{-1}(x)$. Assume also that level sets for the degree are closed and that $M$ is connected; then is it shown in~\cite{VV08} that up to considering some iterate $f^N$ of $f$, we can assume that $\deg_x(f)\geq e^{h_\mathrm{top}(f)}$ for all $x$. The final conditions on the map $f$ are as follows: there exist constants $\sigma>1$ and $L>0$ and an open region $\AAA\subset M$ such that \begin{enumerate}[(H1)] \item $L(x)\leq L$ for every $x\in \AAA$ and $L(x)\leq \sigma^{-1}$ for all $x\in M\setminus \AAA$, and $L$ is close to $1$ (see~\cite{VV08} for precise conditions). \item There exists $k_0\geq 1$ and a covering $\mathcal{P} = \{P_1, \dots, P_{k_0}\}$ of $M$ by domains of injectivity for $f$ such that $\AAA$ can be covered by $r<e^{h_\mathrm{top}(f)}$ elements of $\mathcal{P}$. \end{enumerate} That is, $f$ is uniformly expanding outside of $\AAA$, and does not display too much contraction inside $\AAA$; furthermore, since there are at least $e^{h_\mathrm{top}(f)}$ preimages of any given point $x$, and only $r$ of these can lie in covering of $\AAA$ by elements of $\mathcal{P}$, every point has at least one preimage in the expanding region. The requirement on the potential $\ph$ is as follows: \begin{enumerate}[(P)] \item $\ph\colon M\to \RR$ is H\"older continuous and $\sup\ph - \inf \ph < h_\mathrm{top}(f) - \log r$. \end{enumerate} It is proved in~\cite{VV08} that for any map $f$ and potential $\ph$ satisfying these conditions, there exists a unique equilibrium state for $\ph$. In particular, if (P) holds for $\ph$, then there exists $q_0>1$ such that (P) holds for $q\ph$ as well, for all $q\in (-q_0,q_0)$. Thus Theorem~\ref{thm:birkhoff} applies, and we have the following result on the Birkhoff spectrum. \begin{proposition}\label{prop:VV} Given a map $f\colon M\to M$ satisfying (H1) and (H2) and a H\"older continuous potential $\ph\colon M\to \RR$ satisfying (P), there exists $q_0>1$ such that $T_\BBB$ is $\CCC^1$ on the interval $(-q_0,q_0)$, and writing \[ \alpha_1 = \lim_{q\to -q_0^+} T_\BBB'(q), \qquad \alpha_2 = \lim_{q\to q_0^-} T_\BBB'(q), \] we have $\BBB(\alpha) = \Tb^{L_1}(\alpha)=\inf_{q\in \RR} (T_\BBB(q) - q\alpha)$ for every $\alpha\in [\alpha_1,\alpha_2]$. \end{proposition} See~\cite{VV08} for examples of specific systems to which their conditions, and hence Proposition~\ref{prop:VV}, apply. \subsubsection{Maps with critical points} Ever since the family of logistic maps was introduced, unimodal and multimodal maps have received a great deal of attention. Existence and uniqueness of equilibrium states for a certain class of bounded potentials were established in~\cite{BT08}. In particular, let $\HHH$ denote the collection of topologically mixing $\CCC^\infty$ interval maps $f\colon [0,1]\to[0,1]$ with hyperbolically repelling periodic points and non-flat critical points; given $f\in \HHH$, let $\ph\colon [0,1]\to \RR$ be a H\"older continuous potential such that \begin{equation}\label{eqn:BR} \sup \ph - \inf \ph < h_\mathrm{top} (f). \end{equation} It is shown in~\cite{BT08} that there exists a unique equilibrium state for $\ph$, and so the analogue of Proposition~\ref{prop:VV} holds here. In fact, it was shown by Blokh that any continuous topologically mixing interval map has the specification property (see, for example,~\cite{jB97}), which implies the g-almost product property, and so Pfister and Sullivan's result applies here, showing that the multifractal formalism holds for \emph{any} continuous potential $\ph$ on the entire spectrum. However, their result does not apply to unbounded potentials such as the geometric potential $\ph(x) = -\log \abs{f'(x)}$. The potentials $q\ph$, where $\ph$ is the geometric potential, were studied in~\cite{PS08,BT09,IT09}. In the last of these papers, Iommi and Todd showed that for a related class of maps $f$, the potential $q\ph$ has a unique equilibrium state for all $q\in (-\infty, 0]$. (In fact, they obtain results for $q>0$ as well, but we do not yet have the tools to use these here.) Thus we may apply Theorem~\ref{thm:singularity} and show that if $\alpha_0 = \lim_{q\to 0^-} T_\BBB'(q)$ and $\alpha_\mathrm{min} = \lim_{q\to-\infty} T_\BBB'(q)$, then for all $q\leq 0$, we have \[ T_\BBB(q) = \sup_{\alpha \in \RR} (\BBB(\alpha) + q\alpha), \] and for all $\alpha\in [\alpha_\mathrm{min}, \alpha_0]$, we have \[ \BBB(\alpha) = \inf_{q\in \RR} (T_\BBB(q) - q\alpha). \] In particular, $\BBB(\alpha)$ is strictly concave and $\CCC^1$ on $[\alpha_\mathrm{min},\alpha_0]$, and furthermore, $\Kb{\alpha} = \emptyset$ for $\alpha < \alpha_\mathrm{min}$. \subsection{Entropy spectrum} \subsubsection{Uniform hyperbolicity} For uniformly hyperbolic systems, it can be shown that equilibrium states are Gibbs measures, and so Theorem~\ref{thm:entropy} applies to the entropy spectrum $\EEE(\alpha)$. This gives an alternate proof of a particular case of the results in~\cite{TV99}, where the multifractal analysis of the entropy spectrum is carried out for expansive maps with specification (which includes uniformly hyperbolic systems). \subsubsection{Parabolic maps} Kesseb\"ohmer proves the existence of (non-invariant) weak Gibbs measures for continuous potentials on shift spaces~\cite{mK01}; in~\cite{JR09}, Jordan and Rams examine these weak Gibbs measures as measures on interval maps with parabolic fixed points. Theorem~\ref{thm:entropy} then gives results regarding the entropy spectra of these measures. \subsection{Dimension spectrum} Conformality is automatic for one-dimensional piecewise smooth maps and for rational maps of the Riemann sphere; this provides an ideal setting to apply Theorem~\ref{thm:dimension}. \subsubsection{Uniformly expanding maps} For uniformly expanding maps of the interval, we have $a(x)=|f'(x)|>1$ uniformly, and so $\log a$ is positive and bounded away from $0$. It immediately follows from the remarks in the previous section that all the conditions of Theorem~\ref{thm:dimension} are met. The same results hold on conformal repellers in any dimension, as shown in~\cite{PW97}. Our proof here provides an alternate proof of some of the results in that paper. \subsubsection{Parabolic maps} The dimension spectrum for Manneville--Pomeau maps has been studied in~\cite{kN00,JR09}; once again, the present approach provides an alternate proof of some results. \subsubsection{Maps with critical points} Given a multimodal map $f\in \HHH$, the multifractal analysis of the dimension spectrum for Gibbs measures associated to the potentials described above is carried out in~\cite{mT08,IT09}. At present, these results \emph{cannot} be obtained using the results in this paper, due to the presence of the critical point, which the tools used here cannot yet handle. \section{Preparatory results}\label{sec:prep} \subsection{Convergence results} \begin{proposition}\label{prop:convergence} Let $X$ be a compact metric space, $f\colon X\to X$ be continuous, and $\ph\in \AAA_f$. Let $\mu\in \MMM^f(X)$ be an invariant measure, and consider a sequence of (not necessarily $f$-invariant) measures $\{\mu_n\} \subset \MMM(X)$ such that $\mu_n\to \mu$ in the weak* topology. Then \begin{equation}\label{eqn:Aftoo} \lim_{n\to\infty} \int \ph \,d\mu_n = \int \ph \,d\mu. \end{equation} \end{proposition} \begin{proof} If $\ph$ is continuous, then this is immediate. If $\ph$ is discontinuous, then let $M\in \RR$ be such that $|\ph(x)| \leq M$ for all $x\in X$, and fix $\eps>0$. Condition (B) in the definition of $A_f$ tells us that $\mu(\overline{\CCC(\ph)}) = 0$, and thus there exists an open neighbourhood $B \supset \overline{\CCC(\ph)}$ such that $\mu(\overline{B}) < \eps$. Since $\overline{B}$ is closed, we have \[ \mu(\overline{B}) \geq \ulim_{n\to\infty} \mu_n(\overline{B}), \] and so there exists $N$ such that $\mu_n(B) \leq \mu_n(\overline{B}) < 2\eps$ for all $n\geq N$. Now we have \[ \left\lvert \int_X \ph \,d\mu - \int_X \ph \,d\mu_n \right\rvert \leq \left\lvert \int_{X\setminus B} \ph \,d\mu - \int_{X\setminus B} \ph \,d\mu_n \right\rvert + \left\lvert \int_B \ph \,d\mu - \int_B \ph \,d\mu_n \right\rvert. \] Since $\ph$ is continuous on the compact set $X\setminus B$, the first difference goes to $0$ as $n\to \infty$. Furthermore, by the above estimates, the second difference is less than $3M\eps$. Since $\eps>0$ was arbitrary, this completes the proof of~\eqref{eqn:Aftoo}. \end{proof} \begin{proposition}\label{prop:singularity} Let $X$ be a compact metric space and $\psi\colon X\to \RR\cup\{-\infty\}$ be continuous where finite (and hence bounded above). Consider a sequence of measures $\{\mu_n\}$ converging to $\mu$ in the weak* topology, and suppose that $\int\psi\,d\mu > -\infty$. Then \begin{equation}\label{eqn:sc} \int \psi\,d\mu \geq \ulim_{n\to\infty} \int \psi\,d\mu_n. \end{equation} \end{proposition} \begin{proof} Given $M<0$, define a continuous function $\psi_M\colon X\to \RR$ by \[ \psi_M(x) = \max(\psi(x),M). \] Because $\psi$ is integrable with respect to $\mu$, we have for every $\eps>0$ some $M<0$ such that \[ \int (\psi_M - \psi) \,d\mu < \eps, \] from which we deduce that \[ \int \psi\,d\mu \geq \int \psi_M \,d\mu - \eps = \lim_{n\to\infty} \int \psi_M \,d\mu_n - \eps \geq \ulim_{n\to\infty} \int \psi \,d\mu_n - \eps. \] Because $\eps>0$ was arbitrary, this establishes~\eqref{eqn:sc}. \end{proof} Observe that there are no dynamics in Proposition~\ref{prop:singularity}, so there is no requirement that any of the measures $\mu_n$ or $\mu$ be invariant. \subsection{Measures associated with approximate level sets} Recall that the level sets $\Kb{\alpha}$ are defined by \[ \Kb{\alpha}(\ph) = \left\{ x\in X \,\Big|\, \lim_{n\to\infty} \frac 1n S_n\ph(x) = \alpha \right\}, \] where we write $\Kb{\alpha}(\ph)$ to emphasise the role of the potential function $\ph$. This may be rewritten as \begin{align*} \Kb{\alpha}(\ph) &= \left\{ x\in X \,\Big|\, \forall \eps>0 \exists N \text{ such that } \left| \frac 1n S_n\ph(x) - \alpha \right| < \eps \text{ for all } n\geq N \right\} \\ &= \bigcap_{\eps>0} \bigcup_{N\in\NN} \bigcap_{n\geq N} \left\{ x\in X\,\Big|\, \left| \frac 1n S_n\ph(x) - \alpha \right| < \eps \right\}. \end{align*} In the proofs of our main results, we will need to consider the following ``approximate level sets'': \begin{equation}\label{eqn:Faen} \begin{aligned} {F_\alpha^{\eps,N}}(\ph) &= \bigcap_{n\geq N} \left\{ x\in X\,\Big|\, \left| \frac 1n S_n\ph(x) - \alpha \right| < \eps \right\} \\ {F_\alpha^\eps}(\ph) &= \bigcup_{N\in\NN} {F_\alpha^{\eps,N}}(\ph). \end{aligned} \end{equation} For these we have \[ \Kb{\alpha}(\ph) = \bigcap_{\eps>0} {F_\alpha^\eps}(\ph), \] In particular, the following relations will be quite useful: \begin{align*} h_\mathrm{top} \Kb{\alpha}(\ph) &\leq h_\mathrm{top} {F_\alpha^\eps}(\ph) = \sup_N \left(h_\mathrm{top} {F_\alpha^{\eps,N}}(\ph)\right), \\ \dim_H \Kb{\alpha}(\ph) &\leq \dim_H {F_\alpha^\eps}(\ph) = \sup_N \left(\dim_H {F_\alpha^{\eps,N}}(\ph)\right). \end{align*} Observe that for a continuous function $\ph$, each set $\{x\in X \mid |(1/n)S_n\ph(x) - \alpha|<\eps\}$ is a union of intervals, and ${F_\alpha^{\eps,N}}(\ph)$ is a countable intersection of such sets. When $Z$ is such a set, it is reasonable to approximate $h_\mathrm{top} Z$ with $\underline{Ch}_\mathrm{top} Z$, which gives us an upper bound. A similar upper bound applies when we study the topological pressure. The utility of the capacity quantities (entropy and pressure) for our purposes is in the following lemma, which shows that when we deal with sets like ${F_\alpha^{\eps,N}}$ on which the Birkhoff averages converge \emph{uniformly} to a given range of values, then we can build measures with large free energy and with the expected integrals. \begin{lemma}\label{lem:buildmeasure} Let $X$ be a compact metric space, $f\colon X\to X$ be continuous, and $\psi, \zeta \in \AAA_f$. Fix $Z\subset X$ and let $\beta_1,\beta_2\in [-\infty,\infty]$ be given by \[ \beta_1 = \llim_{n\to\infty} \inf_{x\in Z} \frac 1n S_n \psi(x), \qquad \beta_2 = \ulim_{n\to\infty} \sup_{x\in Z} \frac 1n S_n \psi(x). \] Then for every $\gamma>0$ there exists $\mu\in\MMM^f(X)$ satisfying the following: \begin{align} \label{eqn:stillhere} \int \psi \,d\mu &\in [\beta_1,\beta_2], \\ \label{eqn:varprinc} h(\mu) + \int \zeta \,d\mu &\geq \overline{CP}_Z(\zeta) - \gamma. \end{align} \end{lemma} \begin{proof} The construction of $\mu$ satisfying~\eqref{eqn:varprinc} is given in part 2 of the proof of~\cite[Theorem 9.10]{pW75}, and goes as follows. Choose $\delta>0$ such that \[ \ulim_{n\to\infty} \frac 1n\sum_{x\in E_n} e^{S_n \zeta(x)} > \overline{CP}_Z(\zeta) - \gamma, \] where $E_n$ is a maximal $(n,\delta)$-separated set, and define an atomic measure $\sigma_n$ on $E_n$ by \begin{equation}\label{eqn:sigman} \sigma_n = \frac{ \sum_{y\in E_n} e^{S_n \zeta(y)} \delta_y } { \sum_{z\in E_n} e^{S_n \zeta(z) } }. \end{equation} Define $\mu_n$ by \begin{equation}\label{eqn:mun} \mu_n = \frac 1n \sum_{i=0}^{n-1} \sigma_n \circ f^{-i}, \end{equation} and let $\mu$ be any weak* limit of the sequence $\mu_n$---then $\mu$ is invariant, and the estimate~\eqref{eqn:stillhere} follows from Proposition~\ref{prop:convergence} upon observing that for every $\eps>0$, there exists $N\in \NN$ such that $\int \psi\,d\mu_n \in [\beta_1-\eps, \beta_2+\eps]$ for all $n\geq N$. The estimate~\eqref{eqn:varprinc} is shown in the proof in~\cite{pW75}; although the proof there assumes that $\zeta$ is continuous, this is only used to guarantee the convergence $\int \zeta \,d\mu_{n_j} \to \int \zeta \,d\mu$, which in our case is given by Proposition~\ref{prop:convergence}. \end{proof} The full strength of Lemma~\ref{lem:buildmeasure} is only needed in the proof of Theorem~\ref{thm:dimension} (for the dimension spectrum). For the proof of Theorem~\ref{thm:birkhoff} (for the Birkhoff spectrum), we only need the case $\zeta=0$. In particular, in order to prove Theorems~\ref{thm:high-entropy} and~\ref{thm:singularity}, we only need the following two versions of Lemma~\ref{lem:buildmeasure}. \begin{lemma}\label{lem:high-entropy} Let $X$ be a compact metric space, $f\colon X\to X$ be continuous, and $\ph\colon X\to \RR$ be Borel measurable and bounded above and below. Suppose $Z\subset X$ is such that $\underline{Ch}_\mathrm{top}(Z) > \overline{Ch}_\mathrm{top}(\CCC(\ph))$. Fix $Z\subset X$ and let $\beta_1,\beta_2\in [-\infty,\infty]$ be given by \[ \beta_1 = \llim_{n\to\infty} \inf_{x\in Z} \frac 1n S_n \ph(x), \qquad \beta_2 = \ulim_{n\to\infty} \sup_{x\in Z} \frac 1n S_n \ph(x). \] Then for every $\gamma>0$ there exists $\mu\in\MMM^f(X)$ satisfying the following: \begin{align} \label{eqn:stillhere2} \int \ph \,d\mu &\in (\beta_1 - \gamma,\beta_2 + \gamma), \\ \label{eqn:varprinc2} h(\mu) &\geq \underline{Ch}_\mathrm{top}(Z) - \gamma. \end{align} \end{lemma} \begin{proof} For $n\in \NN$ and $\delta>0$, let $P_n^\delta$ be the maximal cardinality of an $(n,\delta)$-separated subset of $Z$, and recall that \[ \underline{Ch}_\mathrm{top}(Z) = \lim_{\delta\to 0} \llim_{n\to\infty} \frac 1n \log P_n^\delta. \] In particular, decreasing $\gamma$ if necessary, we may choose $\delta>0$ such that \begin{equation}\label{eqn:entropies} \overline{Ch}_\mathrm{top}(\CCC(\ph),\delta) < \underline{Ch}_\mathrm{top}(Z) - \gamma < \llim_{n\to\infty} \frac 1n \log P_n^\delta. \end{equation} Writing $h_0 = \llim_{n\to\infty} \frac 1n\log P_n^\delta$, we choose $\eta>0$ such that $h_0 - \eta > \overline{Ch}_\mathrm{top}(\CCC(\ph),\delta)$. Thus there exists $C>0$ such that for every $m\in \NN$ there exists a set $F_m\subset \CCC(\ph)$ such that $\# F_m \leq Ce^{m(h_0-\eta)}$ and $U_m = \bigcup_{x\in F_m} B(x,m,\delta) \supset \CCC(\ph)$. Observe that $U_m$ is open because $f$ is continuous. Given $n\in \NN$, let $E_n$ be an $(n,\delta)$-separated subset of $Z$ with maximum cardinality $\#E_n = P_n^\delta$. Following the previous proof, consider the measures $\sigma_n$ given by~\eqref{eqn:sigman} with $\zeta=0$: \begin{equation}\label{eqn:sigmanm} \sigma_n = \frac{\sum_{x\in E_n} \delta_x}{ \#E_n}. \end{equation} Now we vary the construction slightly; given $0\leq m<n$, we go $n-m$ steps (not $n$) along each orbit: \begin{equation}\label{eqn:munm} \mu_n^m = \frac 1n \sum_{k=0}^{n-m-1} \sigma_n \circ f^{-k}. \end{equation} That is, $\mu_n^m$ is a convex combination of $\delta$-measures evenly distributed across the first $n - m$ points in each orbit that begins in $E_n$. For every $0\leq k < n - m -1$, consider the set \[ B_n^m(k) = \{x\in E_n \mid f^k \in U_m\} = \bigcup_{z\in F_m} f^{-k}(B(z,m,\delta)) \cap E_n. \] Observe that for every $z\in F_m$ and every pair $x\neq y\in f^{-k}(B(z,m,\delta)) \cap E_n$, we have $d(f^i(x), f^i(y)) < \delta$ for all $n - m \leq i < n$, and since $E_n$ is $(n,\delta)$-separated, it follows that $d(f^i(x),f^i(y))\geq \delta$ for some $0\leq i <n - m$. In particular, $f^{-k}(B(z,m,\delta)) \cap E_n \subset Z$ is $(n - m,\delta)$-separated, and hence has cardinality at most $P_{n-m}^\delta$. It follows that \[ \# B_n^m(k) \leq C e^{m(h_0-\eta)} P_{n-m}^\delta, \] and hence \[ \sigma_n(f^{-k}(U_m)) = \frac{\#B_n^m(k)}{\#E_n} \leq C e^{-\eta m + mh_0} \frac{ P_{n-m}^\delta}{P_n^\delta}. \] This holds for all $0\leq k< n - m$, and hence \begin{equation}\label{eqn:badset} \mu_n^m (U_m) \leq C e^{-\eta m} \frac{ e^{mh_0} P_{n-m}^\delta}{P_n^\delta}. \end{equation} Thus in order to bound $\mu_n^m(U_m)$, we need some control of the ratio $P_{n-m}^\delta/P_n^\delta$. Observe that if $P_n^\delta$ is actually equal to $e^{nh_0}$ for all $n$, then~\eqref{eqn:badset} immediately yields the bound $\mu_n^m (U_m) \leq C e^{-\eta m}$. However, $P_n^\delta$ may not grow as uniformly as we would like, so we must be more careful. Given $m\in \NN$, consider the quantity \[ L(m) = \ulim_{n\to\infty} (\log (P_n^\delta) - \log (P_{n-m}^\delta) - mh_0). \] Suppose $L(m) < 0$. Then there exists $\eps>0$ and $N\in \NN$ such that for all $n\geq N$, we have \[ \log (P_n^\delta) - \log (P_{n-m}^\delta) - mh_0 < -\eps. \] In particular, this gives the following for every $k\in \NN$: \[ \log (P_{N+km}^\delta) < \log (P_N^\delta) + kmh_0 - k\eps. \] Dividing by $km$ and taking the limit as $k\to\infty$, we get \[ h_0 = \llim_{n\to\infty} \frac 1n \log (P_n^\delta) \leq \llim_{k\to\infty} \frac 1{N+km} (P_{N+km}^\delta) < h_0 - \frac \eps m, \] a contradiction. This proves that $L(m) \geq 0$, from which we deduce that for every $m\in \NN$, there exists a sequence $n_j = n_j(m) \to\infty$ such that \[ \llim_{j\to\infty} (\log (P_{n_j}^\delta) - \log (P_{n_j-m}^\delta) - m h_0) \geq 0, \] or equivalently, \begin{equation}\label{eqn:Pnj} \llim_{j\to\infty} \frac {P_{n_j}^\delta}{e^{mh_0} P_{n_j-m}^\delta} \geq 1. \end{equation} In combination with~\eqref{eqn:badset}, this will soon give us the bound we need. As in the proof of Lemma~\ref{lem:buildmeasure}, let $\mu^m$ be a weak* limit point of the sequence $\mu_{n_j}^m$ (by passing to a subsequence if necessary, we assume that $\mu_{n_j}^m \to \mu^m$). Invariance of $\mu^m$ and the entropy estimate~\eqref{eqn:varprinc2} hold just as before, so it only remains to show~\eqref{eqn:stillhere2}. Let $M = \sup_{x\in X} \abs{\ph(x)}$, and choose $m$ large enough so that $C e^{-\eta m} <\gamma/2M$. Carry out the above construction for this value of $m$, and observe that because $U_m$ is open, we have \begin{equation}\label{eqn:smallbad} \mu^m(U_m) \leq \ulim_{n_j\to\infty} \mu_{n_j}^m(U_m) \leq C e^{-\eta m} < \frac \gamma{2M}, \end{equation} where the middle inequality follows from~\eqref{eqn:badset} and~\eqref{eqn:Pnj}. Consequently, \begin{multline} \left\lvert \int_X \ph \,d\mu^m - \int_X \ph \,d\mu_{n_j}^m \right\rvert \leq \\ \left\lvert \int_{X\setminus U_m} \ph \,d\mu^m - \int_{X\setminus U_m} \ph \,d\mu_{n_j}^m \right\rvert + \left\lvert \int_{U_m} \ph \,d\mu^m - \int_{U_m} \ph \,d\mu_{n_j}^m \right\rvert. \end{multline} Since $\ph$ is continuous on the compact set $X\setminus U_m$, the first difference goes to $0$ as $j\to\infty$, and by~\eqref{eqn:smallbad}, the second term is less than $\gamma$; this proves \eqref{eqn:stillhere2} for $\mu^m$. \end{proof} \begin{lemma}\label{lem:singularity} Let $X$ be a compact metric space, let $f\colon X\to X$ be continuous, and let $\psi\colon X\to \RR\cup\{-\infty\}$ be continuous where finite (and hence bounded above). Fix $Z\subset X$ and let $\beta\in \RR$ be given by \[ \beta = \llim_{n\to\infty} \inf_{x\in Z} \frac 1n S_n \psi(x). \] Then for every $\gamma>0$ there exists $\mu\in\MMM^f(X)$ satisfying the following: \begin{align} \label{eqn:stillhere3} \int \psi \,d\mu &\geq \beta, \\ \label{eqn:varprinc3} h(\mu) &\geq \overline{Ch}_\mathrm{top}(Z) - \gamma. \end{align} \end{lemma} \begin{proof} The proof is exactly as in Lemma~\ref{lem:buildmeasure} with the choice $\zeta = 0$, $\eta = \psi$, with Proposition~\ref{prop:singularity} taking the place of Proposition~\ref{prop:convergence}. \end{proof} \section{Proof of Theorem~\ref{thm:birkhoff}} The proof of Theorem~\ref{thm:birkhoff} proceeds in three parts, corresponding to the three parts of the theorem. In the first part, we show that $T_\BBB$ is the Legendre transform of $\BBB$, thus establishing~\eqref{eqn:TisBL}. From this, it immediately follows by standard properties of the Legendre transform that $\Tb^{L_1}$ is the concave hull of $\BBB$; that is, it is the smallest concave function greater than or equal to $\BBB$ at all $\alpha$. Part II of the theorem is an easy consequence of the following proposition. \begin{proposition}\label{prop:empty} Suppose that $\Kb{\alpha}$ is non-empty; that is, there exists $x\in X$ such that $\ph^+(x)=\lim_{n\to\infty}\frac1n S_n\ph(x) = \alpha$. Then $P^*(q\ph)\geq \alpha q$ for all $q\in \RR$. \end{proposition} Once Part I is established, Part III of the theorem is proved via the following series of intermediate results. \begin{proposition}\label{prop:concave} Let $\ph$ be Borel measurable and suppose that $\nu_q$ is an ergodic equilibrium state for $q\ph$. Let $\alpha = \int \ph\,d\nu_q$. Then \begin{equation}\label{eqn:BisTLpt} \BBB(\alpha) \geq \Tb^{L_1}(\alpha). \end{equation} \end{proposition} Note the requirement in Proposition~\ref{prop:concave} that the equilibrium state $\nu_q$ be ergodic. It will often be the case that general arguments will give the existence of \emph{non}-ergodic equilibrium states with $\alpha(\nu_q) = \alpha$, but this is not sufficient for our purposes. \begin{proposition}[Ruelle's formula for the derivative of pressure]\label{prop:ruelle} Let $\psi$ and $\phi$ be Borel measurable functions. If the function \[ q\mapsto P^*(\psi + q\phi) \] is differentiable at $q$, and if in addition $\nu_q$ is an equilibrium state for $\psi + q\phi$, then \begin{equation}\label{eqn:derivative} \frac{d}{dq} P^*(\psi + q\phi) = \int_X \phi \,d \nu_q. \end{equation} \end{proposition} \begin{corollary}\label{cor:alpha-interval} Suppose $T_\BBB$ is continuously differentiable on $(q_1,q_2)$ and $q\ph$ has an equilibrium state $\nu_q$ for each $q\in(q_1,q_2)$. Let $\alpha_1 = D^+T_\BBB(q_1)$ and $\alpha_2 = D^-T_\BBB(q_2)$; then for every $\alpha\in(\alpha_1,\alpha_2)$ there exists $q\in \RR$ such that $q\ph$ has an ergodic equilibrium state $\nu_q$ with $\alpha = \int \ph\,d\nu_q$. \end{corollary} Once these results are established,~\eqref{eqn:BisTL} is a direct consequence of Proposition~\ref{prop:concave} and Corollary~\ref{cor:alpha-interval}. It then follows from basic properties of the Legendre transform that $\BBB = \Tb^{L_1}$ has the same regularity as $T_\BBB$ (except for values of $\alpha$ corresponding to intervals on which $T_\BBB$ is affine). \begin{proof}[Proof of part I] We prove~\eqref{eqn:TisBL} by establishing the following two inequalities: \begin{align} T_\BBB &\leq \BBB^{L_2}, \label{eqn:TleqBL} \\ T_\BBB &\geq \BBB^{L_2}. \label{eqn:BleqTL} \end{align} First we prove~\eqref{eqn:TleqBL}. Recall that \[ T_\BBB(q) = P^* (q\ph) = \sup_{\nu\in\MMM^f_E(X)} \left\{h(\nu) + q \int_X \ph \,d\nu \right\}. \] By Birkhoff's ergodic theorem, every ergodic measure $\nu$ has $\nu(\Kb{\alpha})=1$ for some $\alpha$, and so for $\nu$-almost every $x\in \Kb{\alpha}$ (in particular, for \emph{some $x\in \Kb{\alpha}$}), we have $\int_X \ph\,d\nu = \ph^+(x) = \alpha$. It follows that \begin{align*} T_\BBB(q) &= \sup_{\alpha\in\RR} \left( \sup_{\nu \in\MMM^f_E(\Kb{\alpha})} \left\{h(\nu) + q \int_X \ph \,d\nu \right\} \right) \\ &\leq \sup_{\alpha\in\RR} \left(h_\mathrm{top}(\Kb{\alpha}) + q\alpha\right) = \BBB^{L_2}(q), \end{align*} where the inequality $h(\nu)\leq h_\mathrm{top}(\Kb{\alpha})$ follows from Theorem~A2.1 in~\cite{yP98}. Now we prove the reverse inequality~\eqref{eqn:BleqTL}, by showing that $T_\BBB(q) = P^*(q\ph) \geq \BBB(\alpha) + q\alpha$ for all $q,\alpha\in \RR$. To this end, we fix $\eps>0$ and consider the sets ${F_\alpha^\eps}$, ${F_\alpha^{\eps,N}}$ defined in~\eqref{eqn:Faen}. Applying Lemma~\ref{lem:buildmeasure} with $\zeta = 0$, $\psi = \ph$, $Z = {F_\alpha^{\eps,N}}$, and some $\gamma>0$, we obtain a measure $\mu\in \MMM^f(X)$ with $h(\mu) \geq \underline{Ch}_\mathrm{top}({F_\alpha^{\eps,N}})-\gamma$ and $\int \ph \,d\mu \geq \alpha - \eps$. It follows that \[ P^*(q\ph) \geq h(\mu) + q \int \ph \,d\mu \geq \underline{Ch}_\mathrm{top}({F_\alpha^{\eps,N}}) - \gamma + q\alpha - q\eps, \] and since Lemma~\ref{lem:buildmeasure} can be applied with arbitrarily small $\gamma$, we get \[ P^*(q\ph) \geq h_\mathrm{top}({F_\alpha^{\eps,N}}) + q\alpha - q\eps. \] Taking the supremum over all $N$ yields \[ P^*(q\ph) \geq h_\mathrm{top}({F_\alpha^\eps}) + q\alpha - q\eps \geq h_\mathrm{top}(\Kb{\alpha}) + q\alpha - q\eps, \] and since $\eps>0$ was arbitrary, this implies \[ P^*(q\ph) \geq h_\mathrm{top}(\Kb{\alpha}) + q\alpha. \] This holds for all $q,\alpha\in \RR$, which establishes~\eqref{eqn:BleqTL}. \end{proof} We now proceed to the proof of Part II. \begin{proof}[Proof of Proposition~\ref{prop:empty}] Suppose $\alpha\in \RR$ is such that there exists $x\in \Kb{\alpha}$. Consider the empirical measures \[ \mu_{n,x} = \sum_{i=0}^{n-1} \delta_{f^i(x)}. \] Choose any subsequence $n_k$ such that $\mu_{n_k,x}$ converges in the weak* topology to some $\mu \in \MMM^f(X)$. Then by Proposition~\ref{prop:convergence}, we have $\int \ph\,d\mu = \alpha$, and in particular, \[ P^*(q\ph) \geq h(\mu) + \int q\ph\,d\mu \geq q\int \ph\,d\mu \geq q\alpha \] for every $q\in \RR$. \end{proof} Finally, we prove the string of propositions which implies Part III. \begin{proof}[Proof of Proposition~\ref{prop:concave}] Observe that since $\nu_q$ is ergodic, we have $\nu_q(\Kb{\alpha})=1$, and hence $h(\nu_q) \leq h_\mathrm{top}(\Kb{\alpha})$. Thus \begin{align*} \Tb^{L_1}(\alpha) &= \inf_{q'\in \RR} (T_\BBB(q') - q'\alpha') \\ &\leq T_\BBB(q) - q\alpha' = P^*(q\ph) - q\alpha \\ &= h(\nu_q) + \int_X q\ph \,d\nu_q - q\alpha \\ &\leq h_\mathrm{top}(\Kb{\alpha}) = \BBB(\alpha).\qedhere \end{align*} \end{proof} \begin{proof}[Proof of Proposition~\ref{prop:ruelle}] Write $g(q') = P^*(\psi + q'\phi)$. Then for all $q'\in\RR$, we have \begin{align*} g(q') &= P^*(\psi + q'\phi) \\ &= \sup_\nu \left\{ h(\nu) + \int_X \psi \,d\nu + \int_X q'\phi \,d\nu \right\} \\ &\geq h(\nu_q) + \int_X \psi \,d\nu_q + q' \int_X \phi \,d\nu_q \\ &= P^*(\psi + q\phi) + (q'-q) \int_X \phi \,d\nu_q, \\ &= g(q) + (q'-q) \int_X \phi \,d\nu_q, \end{align*} whence \[ g(q') - g(q) \geq (q'-q) \int_X \phi \,d\nu_q. \] In particular, for $q'>q$, we get \[ \frac{g(q') - g(q)}{q'-q} \geq \int_X \phi \,d\nu_q, \] and hence $g'(q)\geq \int_X \phi \,d\nu_q$ (recall that differentiability of $g$ was one of the hypotheses of the theorem), while for $q'<q$, \[ \frac{g(q') - g(q)}{q'-q} \leq \int_X \phi \,d\nu_q, \] and hence $g'(q)\leq \int_X \phi \,d\nu_q$, which establishes equality. \end{proof} \begin{proof}[Proof of Corollary~\ref{cor:alpha-interval}] Since $T_\BBB'$ is continuous, the Intermediate Value Theorem implies that for every such $\alpha$ there exists $q$ such that $T_\BBB'(q)=\alpha$. Thus applying Proposition~\ref{prop:ruelle} with $\psi=0$ and $\phi=\ph$, we see that any equilibrium state $\nu$ for $q\ph$ has $\nu(q\ph)=\alpha$. Choose some such $\nu$; if $\nu$ is not ergodic, then any element in its ergodic decomposition is also an equilibrium state, and we are done. \end{proof} \section{Proof of Theorems~\ref{thm:phase},~\ref{thm:high-entropy}, and~\ref{thm:singularity}} Given the proof of Theorem~\ref{thm:birkhoff} in the previous section, the proofs of Theorems~\ref{thm:phase},~\ref{thm:high-entropy}, and~\ref{thm:singularity} are relatively straightforward. \begin{proof}[Proof of Theorem~\ref{thm:phase}] Recall that the first two parts of Theorem~\ref{thm:birkhoff} hold without any assumptions on $f$, and thus we already have $T_\BBB = \BBB^{L_2}$. It remains only to show that $\BBB(\alpha) \geq \Tb^{L_1}(\alpha)$ for every $\alpha \in [\alpha_\mathrm{min},\alpha_\mathrm{max}]$, given Condition \textbf{(A)}. Given such an $\alpha$, if there exists $q\in \RR$ such that $T_\BBB'(q)=\alpha$, then the proof of Theorem~\ref{thm:birkhoff} shows that $\BBB(\alpha) = \Tb^{L_1}(\alpha)$. Thus we suppose that no such $q$ exists; in this case, let $q_0=Q(\alpha)$ be the (unique) value of $q$ such that \[ T_\BBB(q) \geq T_\BBB(q_0) + (q-q_0)\alpha \] for all $q\in \RR$. (Equivalently, we have $q_0 = -(\Tb^{L_1})'(\alpha)$.) Applying Theorem~\ref{thm:birkhoff} to the subsystem $X_n$, we see that \[ h_\mathrm{top} (\Kb{\alpha} \cap X_n) = \inf_{q\in \RR} (P_{X_n}^*(q\ph) - q\alpha); \] since $q\mapsto P_{X_n}^*(q\ph)$ is assumed to be differentiable on $\RR$, for every $\alpha\in [\alpha_\mathrm{min},\alpha_\mathrm{max}]$ there exists $q_n \in \RR$ such that $A_n(q_n) = \frac d{dq} P_{X_n}^*(q\ph) |_{q=q_n} = \alpha$. Let $\mu_n$ be an ergodic equilibrium state for $q_n\ph$ on $X_n$; then $\int \ph \,d\mu_n = \alpha$ by Proposition~\ref{prop:ruelle}, and so $\mu_n(\Kb{\alpha}) = 1$. Thus we have \begin{equation}\label{eqn:lowerbound} h_\mathrm{top} \Kb{\alpha} \geq h(\mu_n) = P_{X_n}^*(q_n\ph) - q_n\alpha. \end{equation} It follows from convexity of the pressure function that $q_n\to q_0$ as $n$ goes to $\infty$, and by continuity of the pressure function and Condition \textbf{(A)}, this implies that \[ \lim_{n\to\infty} P_{X_n}^*(q_n\ph) = P^*(q_0\ph), \] which together with~\eqref{eqn:lowerbound} shows that $\BBB(\alpha) \geq T_\BBB(q_0) - q_0 \alpha \geq \Tb^{L_1}(\alpha)$. \end{proof} \begin{proof}[Proof of Theorem~\ref{thm:high-entropy}] The proof of Theorem~\ref{thm:high-entropy} mirrors the proof of Theorem~\ref{thm:birkhoff}; the primary difference is that Lemma~\ref{lem:high-entropy} replaces Lemma~\ref{lem:buildmeasure} in the proof of Part I, where we show~\eqref{eqn:TisBL2}. The proof of Proposition~\ref{prop:empty} does not go through in this setting, and so Part II is weakened from the corresponding statement in Theorem~\ref{thm:birkhoff}. The series of propositions in Part III goes through unchanged, as Proposition~\ref{prop:concave}, Proposition~\ref{prop:ruelle}, and Corollary~\ref{cor:alpha-interval} all hold without regard to continuity of the potential $\ph$. Observe that~\eqref{eqn:TleqBL} holds here as well without modification, since its proof does not require any hypotheses on $\ph$. Thus to prove~\eqref{eqn:TisBL2}, it suffices to establish the following inequality for every $q\in I_Q(h_0)$: \begin{equation}\label{eqn:BleqTL2} T_\BBB(q) \geq \sup_{\alpha\in I_A(h_0)} (\BBB(\alpha) + q\alpha). \end{equation} That is, we show that $T_\BBB(q) = P^*(q\ph) \geq \BBB(\alpha) + q\alpha$ for all $q \in I_Q(h_0)$ and $\alpha \in I_A(h_0)$. Observe that if $\BBB(\alpha) \leq h_0$, then since $\alpha\in I_A(h_0)$ we have $\Tb^{L_1}(\alpha) = \inf_{q\in\RR} (T_\BBB(q) - q\alpha) > h_0 \geq \BBB(\alpha)$, and so in particular $T_\BBB(q) \geq \BBB(\alpha) + q\alpha$ for $q\in I_Q(h_0)$. Thus it remains only to consider the case $\BBB(\alpha) > h_0$. As in the proof of~\eqref{eqn:BleqTL} in Theorem~\ref{thm:birkhoff}, we fix $\eps>0$ and consider the sets ${F_\alpha^\eps}$, ${F_\alpha^{\eps,N}}$ defined in~\eqref{eqn:Faen}. Because $h_0 < \BBB(\alpha) = h_\mathrm{top}\Kb{\alpha} \leq h_\mathrm{top}{F_\alpha^\eps} = \sup_N h_\mathrm{top}{F_\alpha^{\eps,N}}$, we can find $N\in \NN$ such that $h_\mathrm{top}{F_\alpha^{\eps,N}} > h_0$, and then apply Lemma~\ref{lem:high-entropy} with $\psi = \ph$, $Z = {F_\alpha^{\eps,N}}$, and some $\gamma>0$ to obtain a measure $\mu$ with $h(\mu) \geq \underline{Ch}_\mathrm{top}({F_\alpha^{\eps,N}})-\gamma$ and $\int \ph \,d\mu \geq \alpha - \eps - \gamma$. It follows that \[ P^*(q\ph) \geq h(\mu) + q \int \ph \,d\mu \geq \underline{Ch}_\mathrm{top}({F_\alpha^{\eps,N}}) - \gamma + q\alpha - q(\eps + \gamma), \] and since Lemma~\ref{lem:high-entropy} can be applied with arbitrarily small $\gamma$, we get \[ P^*(q\ph) \geq h_\mathrm{top}({F_\alpha^{\eps,N}}) + q\alpha - q\eps. \] Taking the supremum over all such $N$ yields \[ P^*(q\ph) \geq h_\mathrm{top}({F_\alpha^\eps}) + q\alpha - q\eps \geq h_\mathrm{top}(\Kb{\alpha}) + q\alpha - q\eps, \] and since $\eps>0$ was arbitrary, this implies \[ P^*(q\ph) \geq h_\mathrm{top}(\Kb{\alpha}) + q\alpha. \] This holds for all $q\in I_Q(h_0)$ and $\alpha\in I_A(h_0)$, which establishes~\eqref{eqn:BleqTL2}. For Part II of Theorem~\ref{thm:high-entropy}, we observe that if $\BBB(\alpha) > h_0$, then we can apply Lemma~\ref{lem:high-entropy} exactly as above to obtain $T_\BBB(q) \geq \BBB(\alpha) + q\alpha$ for all $q\in \RR$, and hence $\Tb^{L_1}(\alpha) \geq \BBB(\alpha) > h_0$ as well, so $\alpha\in I_A(h_0)$. As remarked above, the propositions in Part III go through unchanged, and we are done. \end{proof} \begin{proof}[Proof of Theorem~\ref{thm:singularity}] The proof of Parts I of Theorem~\ref{thm:singularity} is nearly identical to the proof of Theorem~\ref{thm:high-entropy}, with Lemma~\ref{lem:singularity} replacing Lemma~\ref{lem:high-entropy} in the proof of~\eqref{eqn:TisBL}, and with $(-\infty,0)$ and $(-\infty, \alpha_0]$ replacing $I_Q(h_0)$ and $I_A(h_0)$. Part II of Theorem~\ref{thm:singularity} follows from the observation that Proposition~\ref{prop:empty} \emph{does} apply in this setting as follows: if $\Kb{\alpha}$ is non-empty for some $\alpha\in \RR$, then $P^*(q\ph) \geq \alpha q$ for all $q\leq 0$. The proof only requires replacing Proposition~\ref{prop:convergence} with Proposition~\ref{prop:singularity}. Again,~\eqref{eqn:TleqBL} holds here without modification. Furthermore, for any $q\leq 0$ and $\alpha\in \RR$, we may fix $\eps>0$ and apply Lemma~\ref{lem:singularity} with $\psi = q\ph$, $Z = {F_\alpha^{\eps,N}}$, and some $\gamma>0$ to obtain a measure $\mu\in \MMM^f(X)$ with $h(\mu) \geq \underline{Ch}_\mathrm{top}({F_\alpha^{\eps,N}})-\gamma$ and $\int q\ph \,d\mu \geq q\alpha - q\eps$. It follows that \[ P^*(q\ph) \geq h(\mu) + \int q\ph \,d\mu \geq \underline{Ch}_\mathrm{top}({F_\alpha^{\eps,N}}) - \gamma + q\alpha - q\eps, \] and just as in the proof of Theorem~\ref{thm:birkhoff}, we obtain \[ P^*(q\ph) \geq h_\mathrm{top}(\Kb{\alpha}) + q\alpha. \] This holds for all $q\leq 0$ and $\alpha\in\RR$, which establishes~\eqref{eqn:BleqTL}. Part III is once again just as before. \end{proof} \section{Proof of Theorem~\ref{thm:dimension}}\label{sec:last} As in the proof of Theorem~\ref{thm:birkhoff}, we carry out the proof of Theorem~\ref{thm:dimension} in three parts. First, we show that $T_\DDD$ is the Legendre transform of $\mathcal{D}$, establishing~\eqref{eqn:TisDL}. From this, it immediately follows by standard properties of the Legendre transform that $\Td^{L_3}$ is the concave hull of $\mathcal{D}$. Part II of the theorem is an easy consequence of the following proposition. \begin{proposition}\label{prop:emptyD} Given $\alpha\in \RR$, suppose that $\Kd{\alpha}\cap X'$ is non-empty; that is, there exists $x\in X'$ such that $d_\mu(x) = \alpha$. Then $T_\DDD(q) \geq -\alpha q$ for all $q\in \RR$. Furthermore, if there exists $x\in X'$ such that $d_\mu(x) = +\infty$, then $T_\DDD(q) = +\infty$ for all $q<0$. \end{proposition} Part III of the theorem is once again proved via intermediate results similar in spirit to those in the proof of Theorem~\ref{thm:birkhoff}. \begin{proposition}\label{prop:concaveD} Given $q\in \RR$, let $q_n\to q$ and $t_n \to T_\DDD(q)$ be such that $t_n\leq T_\DDD(q_n)$ for all $n$. Fix $\alpha\in \RR$, and suppose that for all $n\in \NN$, there exists an ergodic equilibrium state $\nu_n$ for $\ph_{q_n,t_n}$ such that $\lambda(\nu_n)>0$ and \begin{equation}\label{eqn:alpha-integralD} \alpha=\frac{-\int \ph_1 \,d\nu_n}{\lambda(\nu_n)}. \end{equation} Then $\DDD(\alpha) \geq \Td^{L_3}(\alpha)$. \end{proposition} \begin{proposition}\label{prop:alpha-intervalD} Given $\eta>0$ and $I_Q=(q_1,q_2)$, suppose that the map $(q,t) \mapsto P^*(\ph_{q,t})$ is continuously differentiable on $R_\eta(I_Q)$, and that $\ph_{q,t}$ has an equilibrium state $\nu_{q,t}$ for every $(q,t)\in R_\eta(I_Q)$. Then for every $\alpha\in(\alpha_2,\alpha_1)=(-D^-T_\DDD(q_2),-D^+T_\DDD(q_1))$ there exists a sequence $(q_n,t_n)\to (q,T_\DDD(q))$ such that each $\ph_{q_n,t_n}$ has an ergodic equilibrium state $\nu_n$ satisfying~\eqref{eqn:alpha-integralD}. \end{proposition} As mentioned after the statement of Theorem~\ref{thm:dimension}, we can do away with the talk of sequences of potentials and measures in Propositions~\ref{prop:concaveD} and~\ref{prop:alpha-intervalD} if each $\ph_q$ has an equilibrium state $\nu_q$ with $\lambda(\nu_q)>0$ and if $T_\DDD$ is $\CCC^r$ on $(q_1,q_2)$. The proof in this case goes just like the proof we carry out below. Before proceeding to the proof itself, we pause to collect pertinent results on the relationship between pointwise dimension, local entropy, and the Lyapunov exponent. Given an ergodic measure $\nu\in \MMM^f_E(X)$, the Lyapunov exponent $\lambda(x) = (\log a)^+(x)$ exists and is constant $\nu$-a.e.\ as a consequence of Birkhoff's ergodic theorem. The analogous result for the local entropy $h_\nu(x)$ was proved by Brin and Katok~\cite{BK83}. The following proposition shows (among other things) that together, these imply exactness of the measure $\nu$ when the map $f$ is conformal. \begin{proposition}\label{prop:localdim} Let $f\colon X\to X$ be continuous and conformal with continuous non-vanishing factor $a(x)$, and fix $\nu\in\MMM^f(X)$. Suppose that the local entropy $h_\nu(x)$ and Lyapunov exponent $\lambda(x)$ both exist at some $x\in X$. If $\lambda(x)>0$, then the pointwise dimension $d_\nu(x)$ also exists, and \begin{equation}\label{eqn:localdim} d_\nu(x) = \lim_{n\to\infty} \frac{-\log \nu(B(x,n,\delta))}{S_n\log a(x)} = \frac{ h_\nu(x) }{\lambda(x)}. \end{equation} If $\lambda(x) = 0$ and $h_\nu(x) > 0$, then $d_\nu(x)$ exists and is equal to $+\infty$. \end{proposition} \begin{proof} Fix $\eps>0$; if $\lambda(x) > 0$, choose $\eps < \lambda(x)$. Since $\lambda(x)$ exists we may apply Lemma~\ref{lem:well-behaved} and obtain $\delta = \delta(\eps)>0$ and $\eta = \eta(x) > 0$ such that~\eqref{eqn:diamball} holds for all $n\in\NN$, and hence writing \begin{equation}\label{eqn:rnsn} r_n = \eta \delta e^{-n(\lambda_n(x) + \eps)}, \qquad s_n = \delta e^{-n(\lambda_n(x)-\eps)}, \end{equation} we have \begin{equation}\label{eqn:nuballs} \nu(B(x,r_n)) \leq \nu(B(x,n,\delta)) \leq \nu(B(x,s_n)). \end{equation} Observe that \begin{equation}\label{eqn:logrn} \log r_n = \log (\eta \delta) - S_n \log a(x) - n\eps, \end{equation} and that furthermore, \begin{equation}\label{eqn:logrnratio} \begin{aligned} \frac{\log r_{n+1}}{\log r_n} &= \frac{\log (\eta \delta) - S_{n+1}\log a(x) - (n+1)\eps}{\log (\eta \delta) - S_n \log a(x) - n\eps} \\ &= 1 - \frac{\eps + \log a(f^n(x))}{\log(\eta \delta) - S_n \log a(x) - n\eps}. \end{aligned} \end{equation} Observe that the numerator is uniformly bounded, and that if $\lambda(x) > 0$, the denominator goes to $-\infty$ by the assumption that $\eps < \lambda(x)$, while if $\lambda(x) = 0$, the denominator goes to $-\infty$ because $\left\lvert \frac 1n S_n \log a(x)\right\rvert < \frac{\eps}2$ for all sufficiently large $n$. It follows that the ratio in~\eqref{eqn:logrnratio} converges to $1$, and a similar result holds for $s_n$. The same argument shows that $r_n \to 0$ for all values of $\lambda(x)$, while $s_n\to 0$ provided $\lambda(x)>0$. For future reference, we point out that everything up to this point also holds if $x\in \BBB$ and $\underline{\lambda}(x) > 0$. Now suppose that $\lambda(x)>0$. It follows that \begin{equation}\label{eqn:ratio2} \lim_{n\to\infty} \frac{-\log r_n}{S_n \log a(x)} = \lim_{n\to\infty} \left(1 + \frac{n\eps - \log (\eta\delta)}{S_n\log a(x)}\right) = 1 + \frac{\eps}{\lambda(x)}. \end{equation} and we see from the first inequality in~\eqref{eqn:nuballs} that \[ \frac{\log \nu(B(x,r_n))}{\log r_n} \left(\frac{-\log r_n}{S_n\log a(x)}\right) \geq \frac{-\log \nu(B(x,n,\delta))}{S_n \log a(x)}, \] where we observe that the quantity on the right is exactly the quantity that appears in~\eqref{eqn:localdim}. Letting $n$ tend to infinity, this yields \begin{equation}\label{eqn:ndimgeq} \llim_{n\to\infty} \frac{\log \nu(B(x,r_n))}{\log r_n} \left(1+\frac{\eps}{\lambda(x)}\right) \geq \frac{h_\nu(x)}{\lambda(x)}. \end{equation} Now given an arbitrary $r>0$, let $n$ be such that $r_n \leq r \leq r_{n-1}$; it follows that \[ \frac{\log \nu(B(x,r))}{\log r} \geq \frac{\log \nu(B(x,r_n))}{\log r_{n-1}} = \frac{\log \nu(B(x,r_n))}{\log r_n} \frac{\log r_n}{\log r_{n-1}}, \] and since $\log r_n / \log r_{n-1} \to 1$, we may let $r$ tend to $0$ to obtain \[ \underline{d}_\nu(x) \left( 1+\frac{\eps}{\lambda(x)} \right) \geq \frac{h_\nu(x)}{\lambda(x)}. \] Since $\eps>0$ was arbitrary, this gives \[ \underline{d}_\nu(x) \geq \frac{h_\nu(x)}{\lambda(x)}. \] Using similar estimates on $s_n$, we obtain the upper bound \[ \overline{d}_\nu(x) \leq \frac{h_\nu(x)}{\lambda(x)}, \] which implies~\eqref{eqn:localdim}. It only remains to consider the case $\lambda(x) = 0$. We first observe that in this case we can choose $N$ sufficiently large that $|S_n\log a(x) - \log (\eta\delta)| < n\eps$ for all $n\geq N$, and hence $0 > \log r_n > -2n\eps$. Then the first inequality in~\eqref{eqn:nuballs} gives \[ \frac{\log \nu(B(x,r_n))}{\log r_n} > -\frac{1}{2n\eps} \log \nu(B(x,n,\delta)), \] and taking the limit as $n\to\infty$ gives \[ \ld_\nu(x) > \frac{h_\nu(x)}{2\eps}, \] just as above. Since $\eps>0$ was arbitrary, we have $d_\nu(x)=+\infty$. \end{proof} The following corollaries of Proposition~\ref{prop:localdim} are easily proved by considering generic points for the measure $\nu$. \begin{corollary}\label{cor:youngs} Let $f\colon X\to X$ be continuous and conformal with continuous non-vanishing factor $a(x)$, and fix $\nu\in \MMM^f(X)$ with $\lambda(\nu)>0$. Then $\dim_H \nu = h(\nu) / \lambda(\nu)$. \end{corollary} \begin{corollary}\label{cor:ptwisedim} Let $f\colon X\to X$ be continuous and conformal with continuous non-vanishing factor $a(x)$, and fix $\mu,\nu\in \MMM^f(X)$. Suppose that $\lambda(\nu)>0$, and let $\alpha\in \RR$ be given by \[ \alpha = \frac{\int h_\mu(x) \,d\nu(x)}{\lambda(\nu)}. \] Then $\nu(\Kd{\alpha}(\mu)) =1$, where $\Kd{\alpha}(\mu)$ is the set of points $x\in X$ for which $d_\mu(x)=\alpha$. \end{corollary} Given a little more information about $X$, we can also say something about measures with zero Lyapunov exponent. \begin{corollary}\label{cor:finitedimh} Let $f\colon X\to X$ be continuous and conformal with continuous non-vanishing factor $a(x)$, and suppose that $\dim_H X<\infty$. Then any $\nu\in \MMM^f(X)$ with $\lambda(\nu)=0$ must have $h(\nu)=0$ as well. \end{corollary} \begin{proof} First suppose that $\nu$ is ergodic and that $h(\nu)>0$. Then by Birkhoff's ergodic theorem and the Brin--Katok entropy formula, there exists a set $Y\subset X$ such that $\nu(Y)=1$ and for every $x\in Y$, we have $\lambda(x) = 0$ and $h_\nu(x) = h(\nu) > 0$. It follows from Proposition~\ref{prop:localdim} that $d_\nu(x) = +\infty$, and hence \[ \dim_H X \geq \dim_H \nu = +\infty, \] which contradicts the assumption in Theorem~\ref{thm:dimension} that $\dim_H X<\infty$. \end{proof} A converse of sorts to Proposition~\ref{prop:localdim} is given by the following, which addresses the case where $d_\mu(x)$ exists even though $h_\mu(x)$ and $\lambda(x)$ may not. We exclude points lying in $\mathbf{Z} = \mathbf{Z}(\mu)$. \begin{proposition}\label{prop:localdim2} Let $f\colon X\to X$ be continuous and conformal with continuous non-vanishing factor $a(x)$, and fix $\mu\in\MMM^f(X)$. Suppose that the pointwise dimension $d_\mu(x)$ exists at some point $x\in X' \cap \mathbf{B}$ and is equal to $\alpha$. Then although the local entropy and Lyapunov exponent may not exist at $x$, the ratio of the pre-limit quantities still converges; in particular, we have \begin{equation}\label{eqn:localdim2} \lim_{n\to\infty} \frac{-\log \mu(B(x,n,\delta))}{S_n\log a(x)} = \alpha = d_\mu(x) \end{equation} whenever $\underline{\lambda}(x) > 0$, and $\alpha=\infty$ if $\underline{\lambda}(x) = 0$. \end{proposition} \begin{proof} We deal first with the case $\underline{\lambda}(x) = 0$. In this case, there exists an increasing sequence $n_k$ such that \[ \frac 1{n_k} S_{n_k} \log a(x) \to 0, \] and since $x\notin \mathbf{Z}$, there exists $\delta_0>0$ such that \[ \gamma(\delta) := \llim_{k\to\infty} -\frac 1{n_k} \log \mu(B(x,n_k,\delta)) > \gamma(\delta_0) > 0 \] for any $0<\delta<\delta_0$. Fix $\eps>0$. Because $x\in \BBB$, we may apply Lemma~\ref{lem:well-behaved} to get $r_n$ as in~\eqref{eqn:rnsn} for which~\eqref{eqn:nuballs} holds for $\mu$, and we have $r_{n_k}\to 0$ just as in the proof of Proposition~\ref{prop:localdim}. In particular, for all sufficiently large $k$,~\eqref{eqn:nuballs} gives \[ \frac{\log \mu(B(x,r_{n_k}))}{\log r_{n_k}} > -\frac{1}{2n_k\eps} \log \mu(B(x,n_k,\delta)), \] and it follows that \[ \alpha = \lim_{k\to\infty} \frac{\log \mu(B(x,r_{n_k}))}{\log r_{n_k}} \geq \frac{\gamma(\delta_0)}{2\eps}. \] Since $\eps>0$ was arbitrary, we see that $\alpha=\infty$. (Observe that since the hypothesis of the proposition tells us that $d_\mu(x)$ exists, it suffices to obtain $\ld_\mu(x)=\infty$, as we do here.) We turn now to the case $\underline{\lambda}(x)>0$. As remarked in the proof of Proposition~\ref{prop:localdim}, the computations at the beginning of that proof are valid here as well; everything up to but not including~\eqref{eqn:ratio2} works in the present setting. \eqref{eqn:ratio2} is replaced by the following inequality: \[ \ulim_{n\to\infty} \frac{-\log r_n}{S_n\log a(x)} \leq 1 + \frac{\eps}{\underline{\lambda}(x)}. \] Thus we have the following in place of~\eqref{eqn:ndimgeq}: \begin{align*} d_\mu(x) \left( 1 + \frac{\eps}{\underline{\lambda}(x)}\right) &= \lim_{n\to\infty} \frac{\log \mu(B(x,r_n))}{\log r_n} \left( 1 + \frac{\eps}{\underline{\lambda}(x)} \right) \\ &\geq \ulim_{n\to\infty} \frac{-\log \mu(B(x,n,\delta))}{S_n\log a(x)}. \end{align*} Similar computations with $s_n$ give \[ d_\mu(x) \left( 1 - \frac{\eps}{\underline{\lambda}(x)} \right) \leq \llim_{n\to\infty}\frac{-\log \mu(B(x,n,\delta))}{S_n\log a(x)}, \] and since $\eps>0$ was arbitrary, this suffices to prove~\eqref{eqn:localdim2}. \end{proof} \begin{proof}[Proof of Theorem~\ref{thm:dimension}] We prove part I of the theorem by establishing the following two inequalities: \begin{align} T_\DDD &\leq \DDD^{L_4}, \label{eqn:TleqDL} \\ T_\DDD &\geq \DDD^{L_4}. \label{eqn:DleqTL} \end{align} We begin by proving~\eqref{eqn:TleqDL}. First, observe that we may have $T_\DDD(q)=+\infty$ for some values of $q$. Suppose that this is the case for some $q\in \RR$; then for any sequence $t_n\to+\infty$, we have $P^*(\ph_{q,t_n}) > 0$ for all $n$, and hence there exists a sequence of ergodic $f$-invariant measures $\nu_n$ such that \begin{equation}\label{eqn:pospress} h(\nu_n) + q\int \ph_1\,d\nu_n - t_n \lambda(\nu_n) > 0. \end{equation} Now there are two possibilities. \emph{Case 1.} $\lambda(\nu_n) > 0$ for all $n$. In this case we obtain \[ \frac{h(\nu_n)}{\lambda(\nu_n)} + q\frac{\int \ph_1\,d\nu_n}{\lambda(\nu_n)} > t_n. \] Applying Corollary~\ref{cor:youngs}, we see that the first term is equal to $\dim_H \nu$; furthermore, Corollary~\ref{cor:ptwisedim} together with the weak Gibbs property of $\mu$ gives $\nu_n(K_{\alpha_n}^\mathcal{D}) = 1$, where $\alpha_n = \int\ph_1\,d\nu_n / \lambda(\nu_n)$. Consequently, we have \[ \mathcal{D}(\alpha_n) + q\alpha_n \geq \dim_H \nu_n + q\alpha_n > t_n, \] and it follows that $\DDD^{L_4}(q) = \sup_{\alpha\in \RR} (\DDD(\alpha) + q\alpha) = +\infty$. \emph{Case 2.} There exists $n$ such that $\lambda(\nu_n) = 0$. Then Corollary~\ref{cor:finitedimh} implies that $h(\nu_n) = 0$ as well, and~\eqref{eqn:pospress} gives us that $q\int \ph_1\,d\nu_n > 0$. If $q\geq 0$, this is impossible, since $\int \ph_1 \,d\nu \leq 0$ for all $\nu\in \MMM^f(X)$. If $q<0$, this implies that $\int \ph_1\,d\nu_n < 0$, and hence $\nu_n(\mathbf{Z}) = 0$. Now for $\nu_n$-a.e.\ $x\in X$, we may apply Proposition~\ref{prop:localdim2} to obtain $d_\mu(x) = +\infty$. It follows that $\nu_n(K_\infty^\mathcal{D}) = 1$ and $K_\infty^\mathcal{D} \cap X' \neq \emptyset$, and we once again have $\DDD^{L_4}(q) = +\infty$. Having dealt with the case where $T_\DDD(q)=+\infty$, we now turn our attention to the case where $T_\DDD(q)$ is finite. Given $t<T_\DDD(q)$, we observe that any measure $\nu$ with $h(\nu) + \int \ph_{q,t}\,d\nu > 0$ must also satisfy $\lambda(\nu) > 0$, otherwise we would have $T_\DDD(q) = +\infty$. It follows that \[ P^*(\ph_{q,t}) = \sup \left\{ h(\nu) + \int \ph_{q,t} \,d\nu \,\Big|\, \nu\in \MMM^f_E(X), \lambda(\nu)>0 \right\}. \] Given $\alpha,\lambda \geq 0$, consider the following set: \[ Z_{\alpha,\lambda} = \{ x\in X \mid \ph_1^+(x) = -\alpha\lambda, \lambda(x) = \lambda \}. \] Every ergodic measure $\nu$ is supported on some $Z_{\alpha,\lambda}$, and so we have \[ 0 < P^*(\ph_{q,t}) = \sup_{\alpha\geq 0} \sup_{\lambda>0} \sup \left\{ h(\nu) + \int \ph_{q,t} \,d\nu \,\Big|\, \nu\in \MMM^f_E(X), \nu(Z_{\alpha,\lambda}) = 1 \right\}. \] It follows that there exists some $\alpha$, $\lambda$, and $\nu$ for which $\nu(Z_{\alpha,\lambda}) =1$ and \[ h(\nu) + q \int \ph_1 \,d\nu - t\lambda(\nu) > 0. \] Applying Corollaries~\ref{cor:youngs} and~\ref{cor:ptwisedim} as before, we see that $\nu(\Kd{\alpha})=1$ and \[ (\dim_H \nu - q \alpha - t) \lambda > 0, \] which immediately yields \[ t < \DDD(\alpha) - q\alpha. \] Since $t<T_\DDD(q)$ was arbitrary, this proves~\eqref{eqn:TleqDL}. In order to show~\eqref{eqn:DleqTL}, we show that \begin{equation}\label{eqn:TgeqD} T_\DDD(q) \geq \DDD(\alpha) - q\alpha \end{equation} for every $q\in \RR$ and $\alpha \in \RR$. (Observe that Proposition~\ref{prop:emptyD} deals with the case $\alpha=\infty$.) Recall from~\eqref{eqn:Td} that \[ T_\DDD(q) = \inf\{t\in \RR\mid P^*(q\ph_1 - t\log a)\geq 0\} = \sup\{t\in \RR\mid P^*(q\ph_1 - t\log a) > 0\}, \] and so to establish~\eqref{eqn:TgeqD} (and hence~\eqref{eqn:DleqTL}), it suffices to show that $P^*(q\ph_1 - t\log a) > 0$ for every $t < \DDD(\alpha) - q\alpha$. To this end, fix $q, t\in \RR$ such that $t + q\alpha < \DDD(\alpha) = \dim_H \Kd{\alpha}$. We will build a measure $\nu$ such that \begin{equation}\label{eqn:largefree} h(\nu) + \int q\ph_1 \,d\nu - t\lambda(\nu) > 0, \end{equation} which will suffice to complete the proof of~\eqref{eqn:DleqTL}, by the above remarks. Observe that since $\dim_H \mathbf{Z} = 0$, we have \[ \dim_H (\Kd{\alpha} \setminus \mathbf{Z}) = \dim_H \Kd{\alpha} > t+q\alpha; \] furthermore, it follows from Proposition~\ref{prop:localdim2} that $\underline{\lambda}(x) > 0$ for every $x\in \Kd{\alpha} \setminus \mathbf{Z}$, and so we may apply~\cite[Theorem 2.1]{vC09a} and obtain \[ P_{\Kd{\alpha} \setminus \mathbf{Z}}(-(t+q\alpha)\log a) > 0, \] where $P_Z$ is the (Carath\'eodory dimension) topological pressure on $Z$. Fix $\gamma>0$ small enough that we have \begin{equation}\label{eqn:gammasmall} P_{\Kd{\alpha} \setminus \mathbf{Z}} (-(t+q\alpha) \log a) - \gamma > \gamma > 0. \end{equation} Now define a family of sets as in~\eqref{eqn:Faen}: for every $\eps>0$ and $n\in \NN$, consider the set \begin{equation}\label{eqn:Gaen} {G_\alpha^{\eps,N}} = \left\{ x\in X \,\Big|\, \abs{\frac {-S_n\ph_1(x)}{S_n \log a(x)} - \alpha} \leq \eps \text{ and $S_n\log a(x) > 0$ for all } n\geq N \right\}. \end{equation} We will also make use of the following sets: \begin{equation}\label{eqn:Gae} {G_\alpha^\eps} = \bigcup_{N\in\NN} {G_\alpha^{\eps,N}}. \end{equation} Applying Proposition~\ref{prop:localdim2} and using the fact that $\mu$ is a weak Gibbs measure for $\ph$, we see that for every $x\in \Kd{\alpha} \setminus \mathbf{Z}$, \[ \alpha = d_\mu(x) = \lim_{n\to\infty} \frac{- S_n\ph_1(x)}{S_n\log a(x)}. \] Since $\underline{\lambda}(x) > 0$ for every $x\in \Kd{\alpha} \setminus \mathbf{Z}$, this implies $\Kd{\alpha} \setminus \mathbf{Z} \subset {G_\alpha^\eps}$ for every $\eps>0$. In particular, this implies that \[ P_{\Kd{\alpha} \setminus \mathbf{Z}}(-(t+q\alpha)\log a) \leq P_{G_\alpha^\eps}(-(t+q\alpha)\log a) =\sup_{N\in \NN} P_{G_\alpha^{\eps,N}}(-(t+q\alpha)\log a), \] and so there exists $N\in \NN$ such that \begin{equation}\label{eqn:gammasmall2} P_{G_\alpha^{\eps,N}}(-(t+q\alpha)\log a) - \gamma > \gamma > 0. \end{equation} Now we can apply the general inequality~\cite[(11.9)]{yP98} to obtain \begin{equation}\label{eqn:gammasmall3} \underline{CP}_{{G_\alpha^{\eps,N}}}(-(t+q\alpha)\log a) - \gamma > \gamma > 0. \end{equation} Let $\psi(x) = \ph_1(x) + (\alpha + \eps)\log a(x)$, and observe that for every $x\in {G_\alpha^{\eps,N}}$ and $n\geq N$, we have \[ \abs{-S_n \ph_1(x) - \alpha S_n \log a(x)} \leq \eps S_n \log a(x), \] which gives \[ -S_n \ph_1(x) \leq (\alpha +\eps) S_n \log a(x), \] and in particular, $S_n \psi(x) \geq 0$. We may now apply Lemma~\ref{lem:buildmeasure} with $\psi = \ph_1 + (\alpha+\eps)\log a$, $\zeta = -(t+q\alpha) \log a$, $Z = {G_\alpha^{\eps,N}}$, and $\gamma$ as before, to obtain a measure $\nu\in \MMM^f(X)$ with the following properties: \begin{align} \label{eqn:int} \int \ph_1 \,d\nu + (\alpha + \eps) \lambda(\nu) &\geq 0, \\ \label{eqn:free} h(\nu) - (t+q\alpha) \lambda(\nu) &\geq \underline{CP}_{{G_\alpha^{\eps,N}}}(-(t+q\alpha)\log a) - \gamma > \gamma > 0. \end{align} If $q\geq 0$, then multiplying~\eqref{eqn:int} by $q$ yields \[ \int q\ph_1 \,d\nu + (q\alpha + q\eps) \lambda(\nu) \geq 0, \] and adding this to~\eqref{eqn:free} yields \[ h(\nu) + \int q\ph_1 \,d\nu - t\lambda(\nu) \geq \gamma - q\eps\lambda(\nu). \] We can choose $\eps>0$ small enough such that $\gamma > q\eps\lambda(\nu)$ for any invariant measure $\nu$, and this establishes~\eqref{eqn:largefree}. For $q\leq 0$, we do a similar computation with $\psi = \ph_1 + (\alpha - \eps)\log a$. \end{proof} We now proceed to the proof of Part II. \begin{proof}[Proof of Proposition~\ref{prop:emptyD}] Suppose there exists $x\in \Kd{\alpha} \setminus \mathbf{Z}$, and let $n_k$ be a subsequence such that the empirical measures $\mu_{x, n_k}$ converge to an invariant measure $\nu$. Then $\lambda(\nu)>0$ (otherwise $\alpha=\infty$ or $x\in \mathbf{Z}$) and $-\int \ph_1 \,d\nu = \alpha \int \log a\,d\nu$ (by Proposition~\ref{prop:localdim2} and weak* convergence). It follows that \begin{align*} P^*(q\ph_1 - t\log a) &\geq h(\nu) + \int q\ph_1 \,d\nu - \int t\log a \,d\nu \\ &\geq -\lambda(\nu) (q\alpha + t) \end{align*} for every $q,t\in \RR$. In particular, if $P^*(\ph_{q,t}) \leq 0$, then $q\alpha + t\geq 0$, hence $t\geq -q\alpha$. This holds for all $t\geq T_\DDD(q)$, and consequently $T_\DDD(q) \geq -q\alpha$ as well. As for the case $\alpha=\infty$, we use the above construction and Corollary~\ref{cor:finitedimh} to obtain $\nu\in \MMM^f(X)$ with $\lambda(\nu) = h(\nu) = 0$. Furthermore, since $x\in X'$, we have $\int \ph_1 \,d\nu < 0$, and it follows immediately that $P^*(\ph_{q,t}) > 0$ for all $q < 0$ and $t\in \RR$, hence $T_\DDD(q) = +\infty$ for all $q<0$. \end{proof} It only remains to prove the propositions implying Part III. \begin{proof}[Proof of Proposition~\ref{prop:concaveD}] It follows from Corollary~\ref{cor:ptwisedim} and the weak Gibbs property of $\mu$ that $\nu_n(\Kd{\alpha}) = 1$ for all $n$. Furthermore, from the assumption that $t_n \leq T_\DDD(q_n)$, we have \[ 0 \leq P^*(\ph_{q_n,t_n}) = h(\nu_n) + q_n\int \ph_1 \,d\nu_n - t_n \lambda(\nu_n) = h(\nu_n) - q_n \alpha \lambda(\nu_n) - t_n \lambda(\nu_n), \] and applying Corollary~\ref{cor:youngs} (using the assumption that $\lambda(\nu_n)>0$) gives \[ \dim_H \nu_n \geq q_n \alpha + t_n. \] Since $\nu_n(\Kd{\alpha}) = 1$, this in turn implies \[ \DDD(\alpha) \geq q_n \alpha + t_n, \] and taking the limit as $n\to\infty$ yields \[ \DDD(\alpha) \geq q\alpha + T_\DDD(q) \geq \Td^{L_3}(\alpha).\qedhere \] \end{proof} \begin{proof}[Proof of Proposition~\ref{prop:alpha-intervalD}] As before, it follows from the finiteness of $T_\DDD(q)$ that $\frac{\di}{\di_t} P^*(\ph_{q,t}) = -\lambda(\nu_{q,t}) < 0$ for all $(q,t)\in R_\eta(I_Q)$, and consequently (assuming $n$ is large enough) we may apply the Implicit Function Theorem to obtain a continuously differentiable function $T_n\colon (q_1,q_2) \to \RR$ such that $(q,T_n(q)) \in R_\eta(I_Q)$ for all $q$, and such that \[ P^*(\ph_{q,T_n(q)}) = \frac 1n. \] Furthermore, we have \[ \lim_{n\to\infty} D^+ T_n(q_1) = D^+ T_\DDD(q_1),\qquad \lim_{n\to\infty} D^- T_n(q_2) = D^- T_\DDD(q_2), \] so for every $\alpha$ as in the statement of the proposition, and for all sufficiently large $n$, we have \[ -D^- T_n(q_2) < \alpha < -D^+ T_n(q_1). \] In particular, by the Intermediate Value Theorem, there exists $q_n$ such that $T_n'(q_n) = -\alpha$. Let $t_n = T_n(q_n)$; then by passing to a subsequence if necessary, we may assume that $(q_n,t_n) \to (q,T_\DDD(q))$ for some $q\in I_Q$. Let $\nu_n$ be an ergodic equilibrium state for $\ph_{q_n,t_n}$; because $P^*(\ph_{q_n,t_n}) >0$ and $T_\DDD(q_n) < \infty$, we have $\lambda(\nu_n) > 0$. Finally, we observe that since $P^*(\ph_{q,t})$ is constant along the curve $(q,T_n(q))$, we have \begin{align*} 0 = \frac{d}{dq} P^*(\ph_{q,T_n(q)})|_{q_n} &= \frac{\di}{\di q} P^*(\ph_{q,t})|_{(q_n,t_n)} + T_n'(q_n) \frac{\di}{\di t} P^*(\ph_{q,t})|_{(q_n,t_n)} \\ &= \int \ph_1 \,d\nu_n + \alpha \lambda(\nu_n), \end{align*} and hence $\nu_n$ satisfies~\eqref{eqn:alpha-integralD}. \end{proof} \bibliographystyle{alpha}
\section{Introduction} Deprotonation of silanol (SiOH) groups\cite{book,iler} at water-silica interfaces is one of the most common and important, yet intriguing, interfacial chemical reactions. Silica (SiO$_2$) is a major component of rocks and lines the channels of many nanofluidic devices.\cite{dekker1,baca,yang} Deprotonation governs dissolution rates,\cite{book} affects lipid binding to silica nanostructures,\cite{baca} creates negative surface charges that can be tuned with moderate changes in solution pH to perform desalination and ion gating,\cite{yang} and may even hinder extraction of positively charged crude oil components\cite{oil} from underground deposits. In particular, the atomic level structural details of deprotonated SiOH groups govern both the overall surface charge density and the binding of ions and molecules to immersed silica surfaces. In this work, we apply {\it ab initio} molecular dynamics\cite{cpmd} (AIMD), which takes into account proton dynamics and hydrogen bond network fluctuations in liquid water essential to acid-base reactions in small molecules\cite{sprik,klein,parrin1,chandler} as well as cooperative hydroxyl hydrogen bonding behavior specific to oxide surfaces,\cite{hass,pore} to investigate the enigmatic SiOH deprotonation equilibrium constant as a function of structural motifs. Measurements of interfacial pK$_{\rm a}$ (defined as $-\log_{10}$ K$_{\rm a}$, where K$_{\rm a}$ is the acid dissociation constant) have been revolutionized by surface-sensitive second harmonic generation (SHG) and sum frequency vibrational spectroscopy (SFVS) techniques.\cite{eisenthal1,shen1} In 1992, Ong {\it et al}.\cite{ong} demonstrated that 19\% of silanol groups on fused silica surfaces exhibit a pK$_{\rm a}$ of 4.5, about the same as vinegar (acetic acid), while 81\% exhibit pK$_{\rm a}$=8.5. SFVS experiments on $\alpha$-quartz reached similar conclusions and further suggested that the low-acidity silanol groups reside in regions with strong water-water hydrogen bonds.\cite{shen05} A titration study on silica gel (amorphous silica)\cite{allen} and X-ray photoelectron spectroscopy measurements on quartz\cite{duval} also independently demonstrated the existence of SiOH groups with pK$_{\rm a}$ between 4 and 5.5. Such qualitative agreement on different forms of silica is expected because liquid water is known to react with even crystalline silica to form an amorphous layer.\cite{iler,shultz} These measurements suggest that the earlier, single pK$_{\rm a} \sim $6.8 reported in amorphous silica titration experiments\cite{schindler} may reflect a composite of two types of SiOH. \begin{figure} \centerline{\hbox{(a)\epsfxsize=1.50in \epsfbox{fig1a.ps} } {\epsfxsize=1.50in \epsfbox{fig1b.ps} (b)}} \centerline{\hbox{(c)\epsfxsize=1.50in \epsfbox{fig1c.ps} } {\epsfxsize=1.50in \epsfbox{fig1d.ps} (d)}} \caption[] {\label{fig1} \noindent (a)-(d) Four types of SiOH groups discussed in this work. (a) Hydrogen-bonded; (b) isolated; (c) Q$^3$; (d) Q$^2$. Oxygen, silicon, and hydrogen atoms are colored in yellow, red, and white, respectively. } \end{figure} The acidities of surface silanol groups have been assigned to different chemical connectivities or inter-silanol hydrogen bonding. In accordance with the literature, we differentiate SiOH groups according to whether they are directly hydrogen bond to other SiOH (``H-bonded,'' Fig.~\ref{fig1}a), or are not so hydrogen-bonded (``isolated,'' Fig.~\ref{fig1}b); whether the 4-coordinated Si atom of the SiOH is part of 3 covalent Si-O-Si- linkages (``Q$^3$,'' Fig.~\ref{fig1}c), or only part of 2 Si-O-Si (``Q$^2$,'' Fig.~\ref{fig1}d). It has been suggested that the ratio of H-bonded to isolated SiOH is about 1 to 4, similar to the relative occurrence of pK$_{\rm a}$=4.5 and 8.5;\cite{ong} thus pK$_{\rm a}$=4.5 has been ascribed to isolated silanol groups.\cite{ong,dong,lorenz,fan1} On the other hand, the Q$^2$:Q$^3$ ratio has also been described as either approximately 1:4 or 4:1, which has prompted assignment of the pK$_{\rm a}$=4.5 SiOH group to either Q$^2$ (Refs.~\onlinecite{mori1,nanosilica}) or Q$^3$ (Refs.~\onlinecite{others,shaw2,rosenholm}). The conflicting estimates of the ratios of silanol groups with different structural motifs\cite{mori1,nanosilica,others,shaw2,rosenholm} likely reflect the difficulty brought about by the tendency of liquid water to react with crystalline silica.\cite{shultz} One indisputable experimental finding is that the silanol surface density is $\sigma_{\rm SiOH}\approx$ 4.6~nm$^{-2}$ on well-soaked amorphous samples.\cite{iler,zhuravlev} On the theoretical side, static geochemical models,\cite{hiemstra,bickmore} which do not account for aqueous phase hydrogen bonding and dynamical proton motion, have also been applied, but they have not yet explained the two observed acidity constants,\cite{ong,shen05,allen} while quantum chemistry or DFT methods with a dielectric continuum treatment of the bulk water environment have been limited to calculating the pK$_{\rm a}$ of small silica fragments.\cite{sahai2,rustad1,sefcik} DFT modeling of amorphous silica slabs have also been considered,\cite{mauri} but the pK$_{\rm a}$ estimates therein often do not treat water explicitly or dynamically. (See Supporting Information (SI) for discussions of the significance of hydrogen-bond network fluctuations and excess proton hopping.) AIMD simulations have successfully reproduced the pK$_{\rm a}$ of molecules in aqueous solution\cite{sprik,klein,parrin1} and should be particularly well-suited for distinguishing {\it relative} SiOH pK$_{\rm a}$ that are 4 pH units apart in different environments, provided we can demonstrate that reproducible pK$_{\rm a}$ for chemically equivalent SiOH's can be predicted. Coupled with static high-level quantum chemistry corrections, they provide the most rigorous predictions for liquid state reactions. As computing power has increased, AIMD modeling of liquid water-material interfaces has become viable,\cite{hass,pore,marx,galli2,mundy,car,angelo} although it remains costly because water dynamics is slower at interfaces.\cite{funel,rossky} Furthermore, investigation of multiple reaction sites and/or crystalline facets is often necessary when dealing with material surfaces. In this work, we study the bimodal acid-base behavior of silanol groups\cite{ong,shen05,allen} by performing AIMD simulations to directly calculate the pK$_{\rm a}$ value. Given the absence of well-defined water-crystalline silica interfaces,\cite{shultz} and the fact that the precise atomic structure of amorphous silica surfaces is unknown, we examine six distinct, representative silanol environments. These include hydroxylated $\beta$-crystobalite (100) (Fig.~\ref{fig2}a), hydroxylated $\beta$-cristobalite (100) with one SiOH removed (Fig.~\ref{fig2}b), reconstructed $\beta$-cristobalite (100) (Fig.~\ref{fig2}c), a molecular system (Fig.~\ref{fig2}e), and two distinct SiOH on reconstructed quartz (0001) (Fig.~\ref{fig2}f). They represent the SiOH motifs proposed to be responsible for pK$_{\rm a}$=4.5 or 8.5 in the literature (Figs.~\ref{fig1}a-d). \begin{figure} \centerline{\hbox{ (a) \hspace*{1.35in} (b) \hspace*{1.35in} (c)}} \centerline{\hbox{\epsfxsize=1.00in \epsfbox{fig2a.ps} } {\epsfxsize=1.00in \epsfbox{fig2b.ps} } {\epsfxsize=1.00in \epsfbox{fig2c.ps} }} \centerline{\hbox{ (d) \hspace*{1.35in} (e) \hspace*{1.35in} (f)}} \centerline{\hbox{\epsfxsize=1.00in \epsfbox{fig2d.ps} } {\epsfxsize=1.00in \epsfbox{fig2e.ps} } {\epsfxsize=1.00in \epsfbox{fig2f.ps} }} \caption[] {\label{fig2} \noindent The heterogeneous SiOH environments examined in this work. (a) Hydroxylated $\beta$-cristobalite (100) surface. The SiOH groups have $\sigma_{\rm SiOH} \sim 8$ nm$^{-2}$, are Q$^2$ and H-bonded. pK$_{\rm a}$=$7.6~\pm 0.3$. (b) Hydroxylated $\beta$-cristobalite (100) surface with one SiOH group replaced with a SiH to break the chain of hydrogen bonds. The tagged (deep blue) SiOH (Q$^2$ and isolated) exhibits pK$_{\rm a}$=8.9$\pm 0.3$. (c) Reconstructed $\beta$-cristobalite (100) surface, $\sigma_{\rm SiOH} \sim 4$ nm$^{-2}$, Q$^3$ and isolated; pK$_{\rm a}$=8.1$\pm 0.5$ (6 layers of water) and~7.0$\pm 0.4$ (4 layers). (d) The structure in panel (c) comes from removing atoms shown here in blue, and attaching the resulting undercoordinated Si and O atoms. Panels (a)-(d) represent $\sim $1.5 simulation cells in the lateral direction. (e) (H$_3$SiO)$_3$SiOH, which is Q$^3$ and isolated, exhibits pK$_{\rm a}$=7.9$\pm 0.5$. (f) Top half of a reconstructed quartz (0001) surface model containing cyclic silica trimers (Si-O)$_3$, $\sigma_{\rm SiOH} \sim 2.3$ nm$^{-2}$, Q$^3$ and H-bonded; one SiOH resides on a trimer ring (pK$_{\rm a}$=5.1$\pm 0.3$), the other does not (pK$_{\rm a}$=3.8$\pm 0.4$ and pK$_{\rm a}$=4.8$\pm 0.4$ depending on whether a nearby trimer ring breaks; see text). Si, O, and H atoms are in yellow, red, and white, respectively. (b) and (e) are finite-temperature AIMD snapshots, with water molecules omitted for clarity, while (a), (c), and (f) are shown at T=0~K. } \end{figure} \section{Computational Methods} AIMD simulations apply the Perdew-Burke-Ernezhof (PBE) functional,\cite{pbe} the Vienna Atomic Simulation Package (VASP),\cite{vasp,vasp1} a 400~eV energy cutoff, $\Gamma$-point sampling of the Brillouin zone, deuterium mass for all protons, and a 0.375~fs time step at each Born-Oppenheimer dynamics time step. The trajectories are thermostat at T=425~K; elevated temperature is needed to represent liquid water properties when the PBE functional is applied and quantum nuclear effects are neglected, which is the case herein (SI, Sec.~S1). Four to six umbrella sampling windows of 20~ps production trajectory length each are used per pK$_{\rm a}$ calculation on hydroxylated $\beta$-cristobalite (100) (Figs.~\ref{fig2}a,b) and reconstructed $\beta$-cristobalite (Fig.~\ref{fig2}c) surfaces. These periodically replicated simulation cells measure 10.17$\times$10.17$\times$26 \AA$^3$. Their lateral dimensions are commensurate with simulation cells often used for AIMD studies of pure water structure or ion hydration. This is one of the reasons we have focused on the (100) rather than the (111) surface of $\beta$-cristobalite, which actually has a $\sigma_{\rm SiOH}$ similar to that of amorphous silica.\cite{iler} The more computationally costly reconstructed quartz (0001) system has a cell size of 10.0$\times$17.32$\times$24 \AA$^3$, and 12 to 18~ps trajectories are used per window. For each crystalline silica simulation cell, we always start from crystal slab structures optimized at zero temperature using Density Functional Theory (DFT) and the PBE functional.\cite{pbe} Next we switch to the CHARMM SiO$_2$ force field\cite{charmm} and the SPC/E model for water.\cite{spce} The number of water molecules occupying the simulation cell is determined using these force fields, the Grand Canonical Monte Carlo (GCMC) technique, and the Towhee code.\cite{gcmc} With this approach, the spaces between hydroxylated and reconstructed $\beta$-cristobalite surfaces are filled with 58 and 63 water molecules, respectively, which amount to roughly six layers of water, while the reconstructed quartz simulation cell contains about 63 (about 4 layers of) water molecules. The (H$_3$SiO)$_3$SiOH pK$_{\rm a}$ calculation utilizes a ($12.42$ \AA)$^3$ cell with 57~H$_2$O and 20~ps sampling trajectories. Finally, to check system size dependences, umbrella sampling simulations for a smaller reconstructed $\beta$-cristobalite (100) simulation cell, 20~\AA\, in the $z$-direction and containing 44 (4 layers of) water molecules, are also conducted for at least 10~ps per window. pK$_{\rm a}$ has been reported for molecules in liquid water using the AIMD technique.\cite{sprik,klein,parrin1} It is related to the standard state deprotonation free energy $\Delta G^{(0)}$ via $-\log_{10}$ $\exp(-\beta \Delta G^{(0)})$, where $\beta$ is $1/k_{\rm B}T$ and \begin{eqnarray} \Delta G^{(0)} = -k_{\rm B}T {\rm ln} \bigg\{ C_0 \int_0^{R_{\rm cut}} dR \, A(R) \, \exp[-\beta \Delta W(R)] \bigg\} \, . \label{eq1} \end{eqnarray} Here $C_0$ denotes 1.0~M concentration, $R$ is the reaction coordinate, $A(R)$ is a phase space factor to be discussed below, $R_{\rm cut}$ is the cutoff distance delimiting the reaction and product valleys in the free energy landscape, and $W(R)$ is the potential of mean force which provides the information needed to compute the free energy of deprotonation. Regardless of the reaction coordinate used,\cite{sprik,klein} $W(R)$ generally do not exhibit turning points in the deprotonated region, and $R_{\rm cut}$ can be taken as the onset of the plateau where $W(R) \rightarrow 0$. The umbrella sampling method\cite{book1} is used to compute the $W(R)$ associated with SiOH deprotonation. A four-atom reaction coordinate $R$ (Fig.~\ref{fig3}a) is found to work best under our simulation conditions. It controls what we call the ``wandering proton'' problem. We label the first, second and third neighbor H$_2$O molecules of the SiO$^-$ oxygen (green sphere) shown in Figs.~\ref{fig3}b-d ``water 1'' (O depicted red), ``2'' (blue), and ``3'' (pink). When $R \sim -0.4$~\AA\ (Fig.~\ref{fig3}b), the SiOH bond is intact. As $R$ decreases to $\sim -1.1$~\AA\ (Fig.~\ref{fig3}c), the SiOH proton is transferred to a ``water 1'' which has been hydrogen-bonded to the SiOH group, yielding a SiO$^-$-H$_3$O$^+$ contact ion pair. As $R$ further decreases (Fig.~\ref{fig3}d), a proton originally residing on ``water 1'' is now transferred to a second water molecule (``water 2''), creating a water-separated SiO$^-$/H$_3$O$^+$ pair, at which point the deprotonation reaction is almost complete. This analysis appears consistent with insights from a transition path sampling AIMD simulation,\cite{chandler} as follows. The Fig.~\ref{fig3}d configuration, with the excess proton and the SiO$^-$ separated by two hydrogen bonds, form a possible free energy dividing surface between the intact and the deprotonated acid species, provided that the electric polarization between the two states, arising from the surrounding water molecules,\cite{chandler} has sufficient time to equilibrate. Our simulation conditions allow such equilibration, and the excess proton is indeed observed to diffuse away if $R$ fluctuates to regions significantly more negative than $\sim -1.4$~\AA. Umbrella potentials of the type $(A/2) (R-R_o)^2$ are used to sample the reaction coordinate $R$. Just as significantly, they ensure that $R>-1.4$~\AA\, and control the extent of proton transfer from ``water 1'' to all possible ``water 2,'' so that at most a water-separated ion-pair is obtained. Otherwise, if the excess proton is several hydrogen bonds removed from the SiO$^-$ (say if it spends a significant amount of time on ``water 3'' via proton transfer from ``water 2''), it can start to diffuse (``wander'') through the simulation cell via the Grotthuss mechanism at O(1)~ps time scale per proton transfer. With tens of H$_2$O molecules in the simulation cell and 10-20~ps trajectories, once the excess H$^+$ leaves the second hydration shell of the SiO$^-$ it does not return, and equilibrium sampling is not achieved. This ``wandering'' likely arises because the higher temperature and longer umbrella sampling trajectories than are generally used in the AIMD literature facilitate diffusion of the excess proton away from the SiO$^-$. Other deprotonation reaction coordinates used in the literature are discussed in the SI (Sec.~S2). They are found to give rise to wandering excess protons under our simulation conditions. As long as equilibrium sampling is achieved, the deprotonation free energy cost should not depend on the choice of coordinate. To apply Eq.~\ref{eq1}, we use a method similar to Ref.~\onlinecite{sprik}: finding the most probable optimal O$_{\rm (water~1)}$-H$^+$ hydrogen bond distance $r_{\rm O-H}$ at each $R$, thus locally converting $W(R)$ to $\bar{W}(r_{\rm O-H})$; performing a spline fit to the resulting $\bar{W}(r_{\rm O-H})$; and integrating over $r_{\rm O-H}$ with a $4\pi r^2_{\rm O-H}$ volume element, which takes the place of the phase space factor $A(R)$ in Eq.~\ref{eq1}. Equation~\ref{eq1} assumes that the entropic factors such as rotations of the reactant and products about the O-H axis are adequately sampled in the AIMD trajectories; otherwise additional constraints and entropic factors are introduced.\cite{klein1,co2} SI Sec.~S2 shows that such constraints do not affect water autoionization free energies, partly because the variation in $r_{\rm O-H}$ needed to complete the deprotonation reaction is relatively small. Furthermore, we always reference predicted silanol pK$_{\rm a}$ to that of water autoionization\cite{klein} computed using the same reaction coordinate and elevated temperature. This minimizes systematic error arising from the simulation protocol (SI Sec.~S2), and phase space contributions approximately cancel out. The metadynamics technique,\cite{meta1,meta2,meta3} a promising and powerful alternative to umbrella sampling, has been applied to calculate dissociation free energies on surfaces\cite{marx} and acid-base reactions of small molecules.\cite{parrin1} This method, not yet implemented in VASP, can potentially be used for efficient comparative study of pK$_{\rm a}$ on other material surfaces after it has been adapted to deal with the wandering proton problem. Gas-phase, high-level {\it ab initio} calculations are performed to check and correct chemical bonding energies predicted with the PBE functional used in the AIMD simulations. The Gaussian03 program suite is applied.\cite{g03} The pertinent sample chemical reaction is \begin{equation} \mathrm{Si(OH)_4 + H_2O \cdots OH^-} \rightarrow \mathrm{Si(OH)_3O^- \cdots H_2O + H_2O} \label{rxn1} \\ \end{equation} Geometries are optimized, and harmonic vibrational frequencies computed, with density functional theory using the B3LYP method\cite{lyp,b3lyp} and the 6-311++G($d,p$) basis set. At the B3LYP geometries, energies are computed with the coupled-cluster singles and doubles method including a perturbative correction for triple substitutions, CCSD(T),\cite{CCSD(T)} using the aug-cc-pVDZ basis set. Basis set incompleteness corrections are added to the CCSD(T) energies. Finally, zero-point vibrational energy corrections computed from the B3LYP/6-311++G($d,p$) frequencies are added. Using this protocol, the high-level {\it ab initio} calculation yields a reaction energy of -30.51 for Eq.~\ref{rxn1}. Gas-phase VASP-based PBE calculations, conducted with an energy cutoff identical to that in AIMD simulations, predict -27.17~kcal/mol. These numbers do not include the 2.24 kcal/mol zero point energy (ZPE) corrections. Thus, the overall AIMD reaction energies should be corrected by a modest -1.10~kcal/mol. The basis set extrapolation procedure may have a systematic error larger than 1 pH unit (SI Sec.~S3) but this does not affect the relative pK$_{\rm a}$ of different SiOH groups. \begin{figure} \centerline{\epsfxsize=3.25in \epsfbox{fig3a.ps}} \centerline{\hbox{\epsfxsize=0.90in \epsfbox{fig3b.ps} \epsfxsize=0.90in \epsfbox{fig3c.ps} \epsfxsize=0.90in \epsfbox{fig3d.ps} }} \centerline{\hbox{ (a) \hspace*{1.35in} (b) \hspace*{1.35in} (c)}} \caption[] {\label{fig3} \noindent (a) The 4-atom reaction coordinate $R$, illustrated for silicic acid in water but is similar for all silanol containing species. Panels (b)-(d) are snapshots from AIMD deprotonation simulations, with outershell H$_2$O molecules removed for clarity reasons. As deprotonation proceeds, $R$ progresses from intact SiO-H ($R \sim -0.4$~\AA, panel (b)) to SiO$^-$ H$_3$O$^+$ contact ion pair ($R \sim -1.0$~\AA, panel (c)) and then, via a Grotthuss proton transfer, to a solvent-separated SiO$^-$/H$_3$O$^+$ pair ($R \sim -1.32$~\AA, panel (d)). Yellow, red, white, and green spheres represent Si, O, H, and the ``O$^-$'' atoms, respectively. Additionally, the water O atoms which are second and third nearest neighbor to the SiO$^-$ oxygen are colored blue and pink, respectively. pK$_{\rm a}$ predictions for silicic acid will be reported in a future publication. } \end{figure} See the SI for details about constraints introduced to prevent proton attacks on SiO$^-$, the AIMD intialization protocol, further justifications for adopting the four-atom reaction coordinate, extrapolating quantum chemistry results to the infinite basis set limit. \section{Results} {\it Hydroxylated $\beta$-cristobalite (100): chemically homogeneous SiOH.} We first show that our simulation protocol predicts reproducible pK$_{\rm a}$ for chemically equivalent SiOH groups on the hydroxylated $\beta$-cristobalite (100) surface (Fig.~\ref{fig2}a). This well-studied model crystalline surface exhibits $\sigma_{\rm SiOH}$=8~nm$^{-2}$, larger than the experimental value of 4.6~nm$^{-2}$ for amorphous silica.\cite{iler} At zero-temperature, it features two types of Q$^2$ silanol groups: alternating hydrogen bond donors and acceptors arranged in chains\cite{meng} (Fig.~\ref{fig2}a) not found in small molecules.\cite{sprik,klein,parrin1} This feature is dynamically preserved in our finite-temperature, aqueous-phase simulations (SI Sec.~S4). Figure~\ref{fig4}a shows that two chemically equivalent, hydrogen bond-accepting SiOH groups on this surface are predicted to exhibit deprotonation $W(R)$ within 0.5~kcal/mol of each other. At large negative values of the reaction coordinate $R$, the deprotonated SiO$^-$ is stabilized with three hydrogen bonds (i.e., $N_w$=3, Figs.~\ref{fig4}b and~\ref{fig4}c). At $R >\sim$ $-$0.8~\AA, the SiO-H bond is only partially broken, $N_w <$~3, and the local $W(R)$ is not sensitive to the slight difference in $N_w$ in the two simulations that arises from statistical noise. These observations appear consistent with a recent two-dimensional potential of mean force analysis of formic acid deprotonation.\cite{parrin1} Accounting for zero-point energy, correcting the AIMD functional with more accurate quantum chemistry methods, and referencing Eq.~\ref{eq1} to the water autodissociation constant pK$_{\rm w}$=14,\cite{klein} we estimate pK$_{\rm a}$ values of 7.5 and 7.7, close to the less acidic pK$_{\rm a}$ value reported by Ong {\it et al.}\cite{ong} The standard deviation is 0.3~pH unit. Multiple deprotonation on this surface is discussed in SI Sec.~S5. {\it Heterogeneity: Isolated, H-bonded, Q$^2$, and Q$^3$ SiOH all exhibit pK$_{\rm a} >$ 7.} We next show that, contrary to previous hypotheses,\cite{ong,lorenz,fan1,mori1,nanosilica,others,shaw2,rosenholm} isolated, H-bonded, Q$^2$, and Q$^3$ silanol groups all exhibit pK$_{\rm a} >$ 7.0. We first create an isolated silanol group by replacing a hydrogen bond-donating SiOH group on the hydroxylated $\beta$-cristobalite (100) surface with a SiH so that its neighboring SiOH group is no longer H-bonded (Fig.~\ref{fig2}b). Figure~\ref{fig5} shows that this isolated SiOH exhibits pK$_{\rm a}$=8.9$\pm 0.3$, and is {\itshape less} acidic by 1.2-1.4 pH unit than when the SiOH hydrogen donor is present (Fig.~\ref{fig2}a). This is entropically reasonable because a hydrogen bond-donating SiOH partner stabilizes the neighboring SiO$^-$ alongside two water molecules, while three water molecules are required for an isolated SiO$^-$. AIMD correctly accounts for this effect because it models H$_2$O and SiOH on the same dynamical footing and because water-water and water-silanol hydrogen bond energies are similar.\cite{meng} Hydroxyls on oxides with more ionic character than SiO$_2$ form stronger hydrogen bonds, and indeed the relative abundance of inter-hydroxyl hydrogen bonding may partially be responsible for the crystal facet-dependent acidity of $\alpha$-Al$_2$O$_3$.\cite{fitts} This will be the subject of future, comparative studies. To compare Q$^2$ and Q$^3$ silanol groups, we reconstruct the hydroxylated $\beta$-cristobalite (100) surface by condensing every other pair of H-bonded SiOH groups into a SiOH and a H$_2$O molecule (Fig.~\ref{fig2}c). This involves the elimination of a hydrogen-bond donating OH group plus the proton on its adjacent, hydrogen-bond accepting SiOH (Fig.~\ref{fig2}d). The resulting undercoordinated Si and O atoms are joined together to form a covalent bond, in the process pulling apart the remaining H-bonded SiOH pairs so they are now isolated from one another. A similar structural motif has been considered in the literature.\cite{chuang} This surface has $\sigma_{\rm SiOH}$=4~nm$^{-2}$, with all SiOH groups being Q$^3$, isolated, and residing on silica rings containing at least 5~Si atoms. Such rings should be unstrained, unlike 3-member (Si-O)$_3$ rings discussed below. The pK$_{\rm a}$ is found to be 8.1$\pm 0.5$ (Fig.~\ref{fig5}). We also consider a (SiH$_3$)$_3$SiOH molecule featuring an isolated, Q$^3$ SiOH group (Fig.~\ref{fig2}e), which exhibits a comparable pK$_{\rm a}$=7.9$\pm 0.5$. Thus, in general, Q$^3$ and Q$^2$ silanol groups do not exhibit pK$_{\rm a}$'s that differ by 4 pH units as previously proposed.\cite{lorenz,fan1,mori1,nanosilica,others,shaw2,rosenholm} Over the range 4~nm$^{-2}$ $\leq \sigma_{\rm SiOH} \leq$ 8~nm$^{-2}$, the precise value of $\sigma_{\rm SiOH}$ has little effect on pK$_{\rm a}$. \begin{figure} \centerline{\hbox{ \epsfxsize=3.50in \epsfbox{fig4ab.ps} }} \centerline{\hbox{ (c) \epsfxsize=1.50in \epsfbox{fig4c.ps} } \epsfxsize=1.50in \epsfbox{fig4d.ps} (d) } \caption[] {\label{fig4} \noindent Deprotonation free energy on hydroxylated $\beta$-cristobalite (100) (Fig.~\ref{fig2}a). (a) Potential of mean force $W(R)$. The most negative $R$ values represent SiO$^-$ while $R >\sim$~$-$0.8~\AA\, indicates intact SiO-H bonds. Red and black lines: two different but chemically equivalent SiOH groups; blue: water autoionization. (b) Mean hydration number, $N_w$, in each umbrella sampling window, defined as the number of protons within 2.5~\AA\, (a typical hydrogen bond distance) of the SiO$^-$ oxygen. (c) Snapshot of the interface between water and hydroxylated $\beta$-cristobalite (100), replicated three times in the lateral direction. (d) Side view of slab model with silanol groups forming hydrogen bond chains, highlighting a Q$^2$, H-bonded SiO$^-$ group in its representative hydrogen bonding environment. Yellow, red, and white spheres depict Si, O, and H atoms, respectively. } \end{figure} To some extent, all our crystalline silica models are nano-slits with thin water slabs confined between the surfaces. As the water content decreases, the dielectric solvation of SiO$^-$ and H$^+$ species should decrease, while intact SiOH groups should be weakly affected. Thus one expects a lower acidity and a higher pK$_{\rm a}$ in strongly nano-confined aqueous media.\cite{yb} To examine confinement effects, a pK$_{\rm a}$ calculation is performed for a smaller reconstructed $\beta$-cristobalite (100) simulation cell, 20~\AA\, in the $z$-direction, containing 4 layers of water. This system actually yields a lower pK$_{\rm a}$=7.0$\pm$0.4---but is lower only by 1.1~pH units (dashed brown line in Fig.~\ref{fig5}). It is possible the unexpected pK$_{\rm a}$ decrease between the 6- and 4-water layer models arises from anomalies in the hydrogen bonding network not apparent from visual inspection of water configurations. The water density is also affected by the confinement.\cite{garofalini2} In the 4-layer model, all water molecules are at most two layers away from the crystalline silica surfaces, and AIMD conducted with GCMC-predicted water content is found to yield a second layer H$_2$O density that is 18\% above 1.0~g/cc. However, this is unlikely to change the dielectric response sufficiently to lower the pK$_{\rm a}$ by 1 unit. Assuming one can apply the Born hydration formula for excess proton hydration in this heterogeneous medium, $\Delta G_{\rm hyd} \sim X (1-1/\epsilon_o)$ where $X\approx -264$~kcal/mol for the proton.\cite{tiss} If the 6-layers of water have $\epsilon_o$=80, the 4-layer system must exhibit $\epsilon_o$=140 to make hydration more favorable by 1~pH unit when in fact confinement generally reduces the dielectric constant of water.\cite{pore} In any case, the discrepancy in pK$_{\rm a}$ is actually within two standard deviations and may simply arise from statistical uncertainties. This test suggests that confinement effects are not large for the slit pore geometry\cite{dekker1} down to about 1~nm slit widths. Therefore our reported pK$_{\rm a}$ for 6-water layer systems should be good approximations of pristine crystalline silica surfaces in contact with bulk liquid water. We also stress that all slab geometries in Figs.~\ref{fig2}a-c have been studied using 6-water layer simulation cells, and their {\it relative} pK$_{\rm a}$ should be mostly free of system size effects. However, two-dimensional confinement in cylindrical amorphous silica nanopores\cite{pore,lorenz,schulten,hartnig,dipole} may have a stronger impact on pK$_{\rm a}$.\cite{yb} {\it High acidity and chemical reactions on strained reconstructed quartz surfaces.} Finally, it seems imperative to demonstrate the possibility of an unusually acidic silanol group. The following ``computational existence proof'' is necessarily more speculative than the conclusions about Q$^2$, Q$^3$, isolated, and H-bonded SiOH pK$_{\rm a}$ discussed above, but it emphasizes the likely role of defected regions when accounting for pK$_{\rm a}$=4.5. Having considered $\sigma_{\rm SiOH}$=8 and~4~nm$^{-2}$, we examine an even lower SiOH surface density. A recent experimental study has attached crystal violet dyes to deprotonated silanol groups. Based on the flat, 120~nm$^2$ surface area of the dye molecule, it is proposed that strong acidity is correlated with local SiOH surface density $\sigma_{\rm SiOH}$ $<$ 0.83~nm$^{-2}$.\cite{dong} In our DFT calculations, the optimal geometry of crystal violet when covalently bonded to (HO)$_3$SiO$^-$ is not flat but is substantially distorted. However, this suggestion of low SiOH surface density being associated with the more acidic SiOH appears consistent with other experimental and theoretical observations discussed below. Since most cuts through crystalline forms of silica yield surfaces with substantial hydroxylation,\cite{nangia} we investigate a reconstructed, completely dehydroxylated quartz (0001) model, featuring (Si-O)$_3$ trimer rings, predicted to be metastable in vacuum.\cite{deleeuw1,deleeuw2} 2.3~SiOH groups per square-nanometer are re-introduced by removing two surface Si atoms and performing further reconstruction and hydroxylation to keep all atoms fully coordinated. This yields two types of silanol groups which are hydrogen-bonded to each other; one member of the pair resides on a cyclic trimer while the other does not (Fig.~\ref{fig2}f). Regions devoid of SiOH and dominated by siloxane (Si-O-Si) bridges are hydrophobic, and this model surface may therefore be consistent with hydroxyl groups in hydrophobic pores reported to be unusually acidic on other oxides.\cite{alumina2} We conduct one deprotonation umbrella sampling simulation of a silanol group residing on a cyclic trimer (Fig.~\ref{fig5}, solid violet curve), and two simulations where the SiOH does not reside on a trimer (dashed and dot-dashed violet). Figure~\ref{fig5} indeed shows that these SiOH groups exhibit pK$_{\rm a}$ $\sim 5.1$$\pm 0.3$, 4.8$\pm 0.4$, and $3.8$$\pm 0.4$, respectively --- close to the experimental value of 4.5. More significantly, their average pK$_{\rm a}$ are separated from the median of all other SiOH groups previously examined in this work by 3.4~pH~units. \begin{figure} \centerline{\epsfxsize=3.50in \epsfbox{fig5.ps}} \caption[] {\label{fig5} \noindent Distribution of pK$_{\rm a}$. Black and red lines: $W(R)$ for SiOH on the hydroxylated $\beta$-cristobalite (100) surface (Fig.~\ref{fig2}a); green: same surface but with hydrogen bond-donating SiOH artificially removed (Fig.~\ref{fig2}b); brown, solid and dashed: reconstructed $\beta$-cristobalite (100) (Fig~\ref{fig2}c), with 6 and 4 layers of water in the simulation cell, respectively; yellow: (H$_3$SiO)$_3$SiOH (Fig.~\ref{fig2}e); violet: reconstructed quartz (0001), (Fig.~\ref{fig2}f), with the SiOH residing on a silica trimer ring (solid) or otherwise (dashed and dot-dashed). Blue: $W(R)$ for water autoionization. } \end{figure} Unlike cyclic silica tetramers or larger Si-O rings, cyclic silica trimers are strained.\cite{book,strain,wallace1} At zero temperature, in the absence of water, the Si atom of the SiOH group residing on a silica trimer ring exhibit Si-O-Si angles of 137.5$^o$, 130.4$^o$, and 132.0$^o$. For the SiOH group not residing on a silica trimer, the angles are 154.9$^o$, 131.8$^o$, and 147.8$^o$. A few of these angles deviate substantially from the ideal, unstrained Si-O-Si value of approximately 145$^o$. This likely accounts for the low pK$_{\rm a}$ computed for these SiOH groups. Other trimer rings not decorated with SiOH are also strained, and the slow spatial decay of their surface strain fields\cite{rickman} may also contribute to the high acidity of SiOH group not residing on them. Within hours in moist air,\cite{book,strain,wallace1} cyclic trimer-containing surfaces are known to incorporate water and break open to reduce strain and increase the local $\sigma_{\rm SiOH}$. At the water-reconstructed quartz interface, during umbrella sampling deprotonation of the SiOH group residing on a 3-member ring (Fig.~\ref{fig2}f), we indeed observe a water molecule forming a transient bond with another Si atom on a Si-O trimer 6~\AA\, away from the tagged SiO$^-$ within picoseconds (Fig.~\ref{fig6}b). The resulting 5-coordinated Si has been observed in simulations\cite{deleeuw2,garofalini1,garofalini} and found to be the intermediate in the trimer ring-opening mechanism on wet silica amorphous surfaces in reactive force field and molecular orbital calculations.\cite{garofalini1,garofalini,lasaga} Our AIMD trajectories show that this mechanism remains operative at explicit liquid water-silica interfaces; proton hopping via the Grotthuss mechanism occurs readily, enabling the H$_2$O adsorbed on the surface Si to lose a proton to bulk water, forming a new SiOH group within $\sim 10$~ps (Fig.~\ref{fig6}c). Then, in 2 of the 4 sampling windows, a Si-O bond on the now 5-coordinated Si breaks to open the OH-incorporated (Si-O)$_3$ ring and irreversibly introduce another new SiOH (Fig.~\ref{fig6}d). Our trajectories thus differ from a recent molecular dynamics study of the liquid water-reconstructed quartz interface, where the adsorbed water molecule, not described by a reactive force field, ultimately desorbs from the 5-coordinated surface Si atom without inducing chemical reactions.\cite{deleeuw2} These irreversible side reactions prevent strict equilibrium sampling needed for $W(R)$ calculations. Fortunately, for the SiOH residing on a cyclic trimer (Fig.~\ref{fig6}), analysis of the pre- and post ring-breaking statistics reveals that the nearby chemical reaction has little effect on its pK$_{\rm a}$. We further analyze trimer ring-opening effects on the pK$_{\rm a}$ of another SiOH group, this one not residing on a surface 3-member ring. A H$_2$O incorporation reaction also occurs in the neighborhood of this tagged SiOH (Fig.~\ref{fig7}). We split the sampling windows into two groups: (A) those without irreversible hydrolysis of a nearby silica ring (Fig.~\ref{fig7}a); and (B) those with trimer ring breaking and formation of two new SiOH groups (Fig.~\ref{fig7}b). Then two complete sets of sampling windows spanning the entire deprotonation pathway are spawned from these seed windows, yielding two pK$_{\rm a}$: case A, pK$_{\rm a}$=3.8 (Fig.~\ref{fig5}, dashed violet curve); and case B, pK$_{\rm a}$=4.8 (Fig.~\ref{fig5}, dot-dashed violet). The results show that H$_2$O incorporation and a single ring-breaking event nearby does appear increase the pK$_{\rm a}$ of the tagged SiOH not residing on a trimer ring. However, the increase is only 1.0~pH unit, almost within statistical uncertainties. In case (A), the three Si-O-Si angles on the silica trimer ring average to 121.1$^o$, 134.0$^o$, and 132.2$^o$ along the trajectory; the first refers to the angle where both Si are below the silica surface (``buried''). In case (B), this angle linking the buried Si atoms relaxes significantly to 140.5$^o$. The second angle, which involves the surface Si and a buried Si, remains almost unchanged at 136.0$^o$. (The third linkage is destroyed during ring opening.) Si atoms which no longer participate in strained Si-O-Si linkages should be more stable against H$_2$O attack. \begin{figure} \centerline{\hbox{ (a) \epsfxsize=1.50in \epsfbox{fig6a.ps} (b) \epsfxsize=1.50in \epsfbox{fig6b.ps} }} \centerline{\hbox{ (c) \epsfxsize=1.50in \epsfbox{fig6c.ps} (d) \epsfxsize=1.50in \epsfbox{fig6d.ps} }} \caption[] {\label{fig6} \noindent Water incorporation onto reconstructed quartz. (a) Initial equilibrated snapshot of SiO$^-$ H$^+$ contact ion pair on the reconstructed quartz (0001) surface where the SiO$^-$ resides on a 3-member ring (truncated in this figure). The tagged SiO$^-$ oxygen is in deep blue; its H-bond SiOH donor is out of the frame. All other Si, O, and H atoms are in yellow, red, and white, respectively. (b) Snapshot after $\sim$ 3~ps. To the left of the SiO$^-$, a water molecule (light blue) has attached itself to a Si atom on another silica trimer ring, which is unhydroxylated. That Si becomes 5-coordinated. (c) This water molecule loses a proton to bulk water; one of the Si-O bonds not on the trimer ring becomes stretched. This is reversible as long as another SiOH is not created. (d) In another sampling window for this SiOH, the trimer ring breaks open instead, forming a new SiOH after extracting a proton from water. } \end{figure} From these considerations, we conclude that, despite interference from H$_2$O incorporation reactions, there remains a statistically significant difference in the pK$_{\rm a}$'s on this surface, and the pK$_{\rm a}$'s ranging from 7.0 to 8.9 for all other silanol groups investigated before (Fig.~\ref{fig5}). This finding suggests that strain, low local silanol surface density, hydrophobicity, and low pK$_a$ are correlated on amorphous silica surfaces. Indeed, atomistic model surfaces with low local $\sigma_{\rm SiOH}$ regions almost always exhibit 3-member rings.\cite{deleeuw1,garofalini1,singer} These regions may be in dynamic equilibrium with solvated silica fragments in solution, constantly being dissolved/hydrolyzed and reconstituted when dissolved fragments re-nucleate on hydroxylated regions.\cite{deleeuw2,criscenti,nangia1,meijer} The dynamic equilibrium has recently been demonstrated in Monte Carlo simulations\cite{nangia1} using a reactive silica force field.\cite{garofalini3} We have only observed one ring-opening reaction on each surface. The limited AIMD trajectory length does not conclusively allow us to predict how many trimer rings persist at the liquid water-amorphous silica interface as a function of time. Therefore we do not definitively assign this structure to the observed pK$_{\rm a}$=4.5 SiOH group, and instead pose it as a challenge to experimental work, including single molecule spectroscopy,\cite{ye} to determine whether they are sufficiently abundant over time to account for the 19\% of all silanol groups shown to exhibit high acidity.\cite{ong,shen05} We also point out that, while cyclic silica trimers are well known to react with moist air, other popular crystalline silica model surfaces should also be hydrolytically unstable.\cite{shultz} Thus, the hydroxylated $\alpha$-quartz (0001) and $\beta$-cristobalite (100) surfaces, with $\sigma_{\rm SiOH}$ $\sim$ 9 and 8~nm$^{-2}$ respectively, are often used as models to study the interface between liquid water and generic silica solids in classical MD simulations. However, if these simulations permit chemical reactions over long enough times, we speculate that some of the SiOH groups on such surfaces may also react with water\cite{shultz,nangia1} in a way to reduce the silanol surface density towards the amorphous silica $\sigma_{\rm SiOH}$=4.6~nm$^{-2}$ observed in experiments.\cite{iler,zhuravlev} \begin{figure} \centerline{\hbox{ (a) \epsfxsize=1.50in \epsfbox{fig7a.ps} (b) \epsfxsize=1.50in \epsfbox{fig7b.ps} }} \caption[] {\label{fig7} \noindent Trimer silica ring and Si-O-Si angles. Configurations are taken from AIMD pK$_{\rm a}$ calculations where the SiOH group involved in the pK$_{\rm a}$ calculation does not reside on a silica trimer ring. That tagged SiOH is off-frame; all H$_2$O molecules are removed for clarity. (a) Snapshot along AIMD trajectory where the 3-member ring remains intact (case A). The 3 Si-O-Si angles within the ring average to 121.1$^o$ (between Si atoms on the trimer ring colored yellow and dark blue), 134.0$^o$ (yellow/pink), and 132.2$^o$ (pink/dark blue). (b) Incorporation of a H$_2$O (its oxygen colored light blue) opens the ring (case B). The surviving Si-O-Si angles are 140.5$^o$ (yellow/dark blue) and 136.0$^o$ (yellow/pink). O and H atoms are represented by red and white spheres respectively. } \end{figure} {\it Discussions} AIMD-based potential of mean force calculations simulations have been demonstrated to yield reproducible pK$_{\rm a}$ for chemically equivalent silanol groups. The statistical uncertainties of our simulation protocol are estimated to be about 0.3-0.5~pH unit, consistent with explicit calculations on two chemically equivalent SiOH. Therefore these simulations should reliably distinguish relative pK$_{\rm a}$ of heterogeneous SiOH groups 4 pH units apart. Resolving hydroxyl pK$_{\rm a}$ on other surfaces may remain a challenge if the acidities are less widely separated. Our pK$_{\rm a}$ calculations suggest that comparative studies between SiO$_2$ and other material surfaces will be extremely interesting. One intriguing candidate surface is a quartz surface densely functionalized with carboxylic acid groups. To our knowledge, this is the only other material surface which clearly exhibits bimodal pK$_{\rm a}$ behavior.\cite{geiger1} Other candidates are the different facets of crystalline alumina, which may feature several pK$_{\rm a}$'s unresolved into distinct, measurable components.\cite{alumina2,alumina3,alumina4} Structural motifs such as chemical connectivity and inter-hydroxyl hydrogen bonding have also been invoked to explain the pK$_a$ in these systems; as mentioned in the text, inter-hydroxyl hydrogen bonding may affect the pK$_{\rm a}$ of the more ionic Al$_2$O$_3$ surfaces more strongly than on any form of silica surfaces. See SI, Sec.~S6, for more details on these material systems. Finally, SFVG spectra\cite{shen05} and time-dependent acid-base phenomena on quartz\cite{geiger2} can also be investigated in the future. \section{Conclusions} In this paper, we have performed AIMD pK$_{\rm a}$ calculations on five representative crystalline silica surfaces plus a molecular system exhibiting different silanol (SiOH) structural motifs. From the results, we have conclusively shown that the more acidic of the two pK$_{\rm a}$ observed in experiments cannot, as previously proposed,\cite{dong,lorenz,fan1,mori1,nanosilica,others,shaw2,rosenholm} be explained by the existence of silanol groups with certain chemical connectivities or inter-silanol hydrogen bonding. In fact, we find pK$_{\rm a}\sim$ 4.5 silanol groups only on strained surfaces with sparse silanol coverage. While our demonstration of the existience of such low pK$_{\rm a}$ SiOH groups is necessarily somewhat speculative, this study highlights the role of defected regions as the most promising candidate to explain the elusive bimodal acid-base behavior of silica surfaces.\cite{ong,shen05,allen} Assigning structural motifs to the more strongly acidic SiOH groups is particularly crucial in non-reactive force field-based modeling of silica nanofluidic channels, where the preferentially deprotonated SiOH sites at neutral pH have to be assigned in a static way.\cite{pore,lorenz,hartnig,schulten,schulten1,jctn} In the process of studying the acid-base behavior, we also observe irreversible, water-assisted ring-opening reactions of strained silica trimer rings in contact with liquid water, The reaction was previously studied on wet silica surfaces;\cite{garofalini1,lasaga,garofalini1} our AIMD simulations demonstrate that a similar mechanism is operative at liquid water-silica interfaces. \section*{Acknowledgement} We thank Ron Shen, Steven Garofalini, Susan Rempe, Jeff Brinker, Dave Tallant, Ying-Bing Jiang, and Franz Geiger for discussions. This work was supported by the Department of Energy under Contract DE-AC04-94AL85000. Sandia is a multiprogram laboratory operated by Sandia Corporation, a Lockheed Martin Company, for the U.S.~Deparment of Energy. LJC acknowledge support from the U.S. DOE Office of Basic Energy Sciences, Division of Chemical Sciences, Geosciences, and Biosciences. \section*{Supporting Information Available} Further information are provided regarding details of the AIMD simulations, justification for the reaction coordinate used, quantum chemistry calculations (including all energies and optimized geometries), the dynamics of hydrogen bond fluctuations on silica surfaces, proton exchange due to multiple deprotonation, and a brief overview of acid-base behavior in alumina and carboxylate acid functionalized surfaces. This information is available free of charge via the Internet at {\tt http://pubs.acs.org/}.
\section{ Introduction} Quantum enveloping algebras are one of the two main examples of quantum groups introduced by Drinfeld \cite{Dr1} and Jimbo \cite{Jb} in their study of the Yang-Baxter equation. In the most general definition the role of the quantum Yang-Baxter equation was given in the form of universal $R$-matrix. The first example was defined with the help of Yang-Baxter equation for a $2\times 2$ R-matrix. In the physics literature and also in the work of Reshetikhin \cite{R} the dependence on parameters can be more than one variable (see also \cite{T, J2}). The research on quantum enveloping algebras with two parameters have been revitalized by Benkart and Witherspoon in their work on Drinfeld double construction and Schur-Weyl duality \cite{BW1, BW2, BW3}. For a review of early history the reader to referred to the introduction in \cite{BW1}. After then generalizations to other simple Lie algebras were given by Bergeron, Gao and Hu in \cite{BGH1, BGH2, BH} and their representations are studied accordingly. All these work show that the two parameter quantum groups have similar properties like the usual quantum groups but offer distinct features in two parameter cases. Recently by generalizing the vertex representations \cite{FJ, J1} of quantum affine algebras, Hu, Rosso and the second author \cite{HRZ} have further introduced a two parameter quantum affine algebra for the affine type $A$ and also obtained its Drinfeld realization. It is well-known that quantum affine algebras also admit a fermionic realization \cite{H}. The fermionic realization has played a fundamental role in quantum integrable systems and figured prominently in Kyoto school's work on quantum affine algebras and their applications to statistical mechanics \cite{DJKMO, JM}. The crystal basis \cite{Ka1, Ka2} for the quantum general affine algebra was first constructed with help of Hayashi's fermionic representation \cite{H}. Our first motivation is to generalize this construction to construct all level one fundamental representations for two parameter quantum affine algebra $U_{r,s}(\widehat{\mathfrak {sl}_n})$. In particular, the fermionic realization is also obtained, which is two parameter generalization of Misra-Miwa's realization \cite{MM} and a generalization of Leclerc and Thibon's version of the combinatorical representation \cite{LT}. Our second motivation is somewhat more fundamental to justify the study of two-parametric quantum groups. In the early days the true meaning of various parameters in quantum groups puzzled some researchers to question whether the introduction of other parameters is really necessary. In this paper we will explain the meaning of two parameters and obtain their combinatorial interpretation, and show how nicely many pieces of two-parameter quantum groups are patched together in our fermionic model. Roughly speaking, the two parameters correspond naturally to the left and right (multiplication) in our combinatorial model, and one further sees that the Serre relations are consequence of some combinatorial properties of our model. \section{Quantum Affine Algebra $U_{r,s}(\widehat{\mathfrak {sl}_n})$} \medskip In this section, we will recall the structure of two-parameter quantum affine algebra $U_{r,s}(\widehat{\mathfrak {sl}_n})$ defined in \cite{HRZ}. Let $\mathbb{K}=\mathbb{Q}(r,s)$ denote the field of rational functions in two variables $r$, $s$ ($r\ne \pm s$). For $n\geq 2$ let $\varepsilon_1, \varepsilon_2, \cdots, \varepsilon_{n-1}$ be an orthonormal basis of $E=\mathbb{R}^n$ under the inner product $( \, , )$. Set $I=\{1,\cdots,n-1\}$, $I_0=\{0\}\cup I$. Then $\Phi=\{\varepsilon_i-\varepsilon_j\mid i \neq j\in I\}$ is the set of roots for the simple Lie algebra $\mathfrak{sl}(n)$. We can take $\Pi=\{\alpha_i=\varepsilon_i-\varepsilon_{i+1}\mid i\in I \}$ as the basis of simple roots. Let $\delta$ denote the primitive imaginary root of $\widehat{\mathfrak {sl}_n}$. Take $\alpha_0=\delta-(\varepsilon_1-\varepsilon_n)$, then $\Pi'=\{\alpha_i\mid i\in I_0\}$ is a base of simple roots of affine Lie algebra ${\widehat{\mathfrak{sl}_n}}$. The following definition is an affinization of the two-parameter quantum groups for type $\mathfrak{sl}_n$ (see \cite{BW1}). \begin{defi} Let $U=U_{r,s}(\widehat{\mathfrak {sl}_n})$ $(n\geq 2)$ be the unital associative algebra over $\mathbb{K}$ generated by the elements $e_j,\, f_j,\, \omega_j^{\pm 1},\, \omega_j'^{\,\pm 1}\, (j\in I_0)$, $\gamma^{\pm\frac{1}2}$, $\gamma'^{\pm\frac{1}2}$, $ D^{\pm1}$, $D'^{\,\pm1}$, satisfying the following relations: \medskip \noindent $(\textrm{A1})$ \ $\gamma^{\pm\frac{1}2},\,\gamma'^{\pm\frac{1}2}$ are central with $\gamma=\omega_\delta$, $\gamma'=\omega'_\delta$, such that $\omega_i\,\omega_i^{-1}=\omega_i'\,\omega_i'^{\,-1}=1 =DD^{-1}=D'D'^{-1}$, and \begin{equation*} \begin{split}[\,\omega_i^{\pm 1},\omega_j^{\,\pm 1}\,]&=[\,\omega_i^{\pm1}, D^{\pm1}\,]=[\,\omega_j'^{\,\pm1}, D^{\pm1}\,] =[\,\omega_i^{\pm1}, D'^{\pm1}\,]=0\\ &=[\,\omega_i^{\pm 1},\omega_j'^{\,\pm 1}\,]=[\,\omega_j'^{\,\pm1}, D'^{\pm1}\,]=[D'^{\,\pm1}, D^{\pm1}]=[\,\omega_i'^{\pm 1},\omega_j'^{\,\pm 1}\,]. \end{split} \end{equation*} $(\textrm{A2})$ \ For $\,i,\, j\in I_0$, \begin{equation*} \begin{array}{lll} & D\,e_i\,D^{-1}=r^{\delta_{0i}}\,e_i,\qquad\qquad\qquad\qquad\; &D\,f_i\,D^{-1}=r^{-\delta_{0i}}\,f_i,\\ &\omega_j\,e_i\,\omega_j^{\,-1}=\langle \omega'_i,\,\omega_j\rangle \,e_i,\qquad\quad &\omega_j\,f_i\,\omega_j^{\,-1}=\langle \omega'_i,\,\omega_j\rangle^{-1}\,f_i. \end{array} \end{equation*}\\ $(\textrm{A3})$ \ For $\,i,\, j\in I_0$, \begin{equation*} \begin{array}{lll} & D'\,e_i\,D'^{-1}=s^{\delta_{0i}}\,e_i,\qquad\qquad\qquad\quad\ \ &D'\,f_i\,D'^{-1}=s^{-\delta_{0i}}\,f_i,\\ &\omega'_j\,e_i\,\omega'^{\,-1}_j=\langle \omega'_j,\,\omega_i\rangle^{-1} \,e_i, \qquad\ \ &\omega'_j\,f_i\,\omega'^{\,-1}_j=\langle \omega'_j,\,\omega_i\rangle\,f_i. \end{array} \end{equation*}\\ $(\textrm{A4})$ \ For $\,i,\, j\in I_0$, we have $$[\,e_i, f_j\,]=\frac{\delta_{ij}}{r-s}(\omega_i-\omega'_i).$$ $(\textrm{A5})$ \ For $\,i,\,j\in I_0$, but $(i,j)\notin \{\,(0,n-1), \ (n-1,0)\,\}$ with $a_{ij}=0$, we have $$[\,e_i, e_j\,]=0=[\,f_i, f_j\,].$$ $(\textrm{A6})$ \ For $\,i\in I_0$, we have the $(r,s)$-Serre relations: \begin{gather*} e_i^2e_{i+1}-(r+s)\,e_ie_{i+1}e_i+(rs)\,e_{i+1}e_i^2=0,\\ e_ie_{i+1}^2-(r+s)\,e_{i+1}e_ie_{i+1}+(rs)\,e_{i+1}^2e_i=0,\\ e_{n-1}^2e_0-(r+s)\,e_{n-1}e_0e_{n-1}+(rs)\,e_0e_{n-1}^2=0,\\ e_{n-1}e_0^2-(r+s)\,e_0e_{n-1}e_0+(rs)\,e_0^2e_{n-1}=0. \end{gather*} $(\textrm{A7})$ \ For $\,i\in I_0$, we have the $(r,s)$-Serre relations: \begin{gather*} f_i^2f_{i+1}-(r^{-1}+s^{-1})\,f_if_{i+1}f_i+(r^{-1}s^{-1})\,f_{i+1}f_i^2=0,\\ f_if_{i+1}^2-(r^{-1}+s^{-1})\,f_{i+1}f_if_{i+1}+(r^{-1}s^{-1})\,f_{i+1}^2f_i=0,\\ f_{n-1}^2f_0-(r^{-1}+s^{-1})\,f_{n-1}f_0f_{n-1}+(r^{-1}s^{-1})\,f_0f_{n-1}^2=0,\\ f_{n-1}f_0^2-(r^{-1}+s^{-1})\,f_0f_{n-1}f_0+(r^{-1}s^{-1})\,f_0^2f_{n-1}=0, \end{gather*} where $\langle \omega'_i,\,\omega_j\rangle$ is a skew-dual pairing defined as follows (more detail see \cite{HRZ}): $$\langle \omega'_i,\,\omega_j\rangle =\begin{cases}r^{(\varepsilon_j,\, \alpha_i)}\,s^{(\varepsilon_{j+1},\, \alpha_i)}, \quad\ \; (i\in I_0,\, j\in I)\cr r^{-(\varepsilon_{i+1},\, \alpha_0)}\,s^{(\varepsilon_1,\, \alpha_i)}, \quad (i\in I_0,\,j=0)\end{cases}$$ From now on, let us write briefly $\langle \omega'_i,\,\omega_j\rangle=\langle i,\,j\rangle$. \end{defi} It can be proved (see \cite{HRZ}) that $U_{r,s}(\widehat{\mathfrak {sl}_n})$ is a Hopf algebra with the coproduct $\Delta$, the counit $\varepsilon$ and the antipode $S$ defined below: for $i\in I_0$, we have \begin{gather*} \Delta(\gamma^{\pm\frac{1}2})=\gamma^{\pm\frac{1}2}\otimes \gamma^{\pm\frac{1}2}, \qquad \Delta((\gamma')^{\,\pm\frac{1}2})=(\gamma')^{\,\pm\frac{1}2}\otimes (\gamma')^{\,\pm\frac{1}2}, \\ \Delta(D^{\pm1})=D^{\pm1}\otimes D^{\pm1},\qquad \Delta(D'^{\,\pm1})=D'^{\,\pm1}\otimes D'^{\,\pm1},\\ \Delta(w_i)=w_i\otimes w_i, \qquad \Delta(w_i')=w_i'\otimes w_i',\\ \Delta(e_i)=e_i\otimes 1+w_i\otimes e_i, \qquad \Delta(f_i)=f_i\otimes w_i'+1\otimes f_i,\\ \varepsilon(e_i)=\varepsilon(f_i)=0,\quad \varepsilon(\gamma^{\pm\frac{1}2}) =\varepsilon((\gamma')^{\,\pm\frac{1}2})=\varepsilon(D^{\pm1})=\varepsilon(D'^{\,\pm1})=\varepsilon(w_i)=\varepsilon(w_i')=1, \\ S(\gamma^{\pm\frac{1}2})=\gamma^{\mp\frac{1}2},\qquad S((\gamma')^{\pm\frac{1}2})=(\gamma')^{\mp\frac{1}2},\qquad S(D^{\pm1})=D^{\mp1},\qquad S(D'^{\,\pm1})=D'^{\,\mp1},\\ S(e_i)=-w_i^{-1}e_i,\qquad S(f_i)=-f_i\,w_i'^{-1},\qquad S(w_i)=w_i^{-1}, \qquad S(w_i')=w_i'^{-1}. \end{gather*} \begin{remark} The algebra $U_{r,s}(\widehat{\mathfrak {sl}_2})$ is isomorphic to the quantum affine algebra $U_{q, q^{-1}}(\widehat{\mathfrak {sl}_2})$ if set $rs^{-1}=q^2$, see \cite{HRZ}. \end{remark} \begin{remark} We remark that the two sets of generators $\omega_i, \omega_i'$ follow the original idea of \cite{BW1, BW2} to naturally blend the second parameter into the relations. Roughly speaking when one identifies $\omega_i'$ to $\omega_i^{-1}$ the algebra specializes to the usual quantum affine algebra. \end{remark} \section{Fock space representations of $U_{r,s}(\widehat{\mathfrak {sl}_n})$} In this section we construct a Fock space representation for the quantum affine algebra $U_{r,s}(\widehat{\mathfrak {sl}_n})$ based on the fermionic representation of the usual quantum affine algebra. The Fock space is modeled on the space of partitions. A partition is a decomposition of a natural number written in nondecreasing order. Let $\mathcal P(n)$ be the set of partitions of $n$. The generating function of partitions is given by $$ \sum_{n=0}^{\infty}|\mathcal P(n)|q^n=\prod_{n=1}^{\infty}(1-q^n)^{-1}. $$ For each partition $\lambda=(\lambda_1, \lambda_2, \cdots, \lambda_l)$ we associate the Young diagram consisting of $n$ nodes (or boxes) which are stacked in $l$ rows in the 4th quarter of the xy-coordinate system and aligned at the origin in such a way that the ith row occupies $\lambda_i$ nodes. The diagonal of the Young diagram is given by the nodes along the line $y=-x$. Another way to identify the Young diagram is the following: we specify the path starting at $(0, -\lambda_1')$ and we move eastward by $\lambda_l$ steps and then we go north by $\lambda_1'-\lambda_2'$ steps and so on. We say that a node or a box in $\lambda$ sitting at position $(a, -b)$ if its upper left corner is situated at the point $(a, -b)$. We will define the residue of any node at $(a, -b)$ to be $a-b\,( mod \, n)$. For instance the Young diagram $\lambda=(6,4,4,2,2)$ with residues is given in Figure \ref{F:young}. For convenience we allow the residues to take values in $\mathbb Z_n$. So in Figure 1 the residues shown in the lower left part will be $5, 4, 3$ respectively for the quantum algebra $U_{r,s}(\widehat{\mathfrak {sl}_6})$. \begin{figure}[h] \label{F:young} \setlength{\unitlength}{0.75cm} \begin{picture}(7,5) \multiput(0,3)(1,0){7}% {\line(0,1){1}} \multiput(0,3)(0,1){2}% {\line(1,0){6}} \multiput(0,2)(1,0){5}% {\line(0,1){1}} \multiput(0,2)(0,1){1}% {\line(1,0){4}} \multiput(0,1)(1,0){3}% {\line(0,1){1}} \multiput(0,1)(0,1){1}% {\line(1,0){2}} % \multiput(0,0)(1,0){3}% {\line(0,1){1}} \multiput(0,0)(0,1){1}% {\line(1,0){2}} \multiput(0.4,3.3)(1, -1){2}{\mbox{$0$}} \multiput(1.4,3.3)(1, -1){2}{\mbox{$1$}} \multiput(2.4,3.3)(1,-1){2}{\mbox{$2$}} \multiput(3.4,3.3)(1,0){1}{\mbox{$3$}} \multiput(4.4,3.3)(1,0){1}{\mbox{$4$}}\put(5.4,3.3){\mbox{$5$}} \multiput(0.1, 2.3)(1,-1){2}{\mbox{${-1}$}} \multiput(0.1, 1.3)(1,-1){2}{\mbox{$-2$}} \put(0.1, 0.3){\mbox{$-3$}} \end{picture}\par \caption{Young diagram $(64^22^2)$} \end{figure} A node $\gamma$ or box (convex corner) is called {\it removable} if $\lambda-\gamma$ is still a Young diagram. A contract corner is called {\it intent} if one node or box can be added at the corner to form another Young diagram. By abusing the terminology we will call a node $(a, b)$ of the border rim of the diagram of $\lambda$ {\it indent} if the box $(a-1, b-1)$, $(a+1, b)$ or $(a, b-1)$ can be added to $\lambda$ to form a new Young diagram. Note that the new diagram's border does not contain the node except the node is a starting node or a final node of the border. For example in Figure 1 the nodes at $(1, -1)$, $(3, 0), (5, 0), (0, -3)$ are indent nodes except that the nodes at $(5, 0)$, $(3, -1)$, $(1, -3)$ are removable nodes. By convention the trivial Young diagram $\phi$ has an imaginary indent node at $(-1, 1)$. For this reason we sometimes say that a point at $(a, -b)$ is removable or indented. In this sense the origin for the trivial Young diagram is indented. For each $i\in \mathbb Z_n$ we define the Young diagram $|\lambda, i\rangle$ as the diagram that assigns the residues $i+a-b \, ( mod \, n)$ to each node $(a, -b)$ in $\lambda$. The previous example is the Young diagram for $i=0$. The following Figure 2 shows the residue of $(a, -b)$ for the configuration of $|\lambda, i\rangle$. \begin{figure}[h] \label{F:young2} \setlength{\unitlength}{0.75cm} \begin{picture}(7,7) \put(0, 5){\line(1,0){7}} \put(0,0){\line(0,1){5}} \put(7,4){\line(0,1){1}}\put(7, 4){\line(-1,0){1}} \put(6,4){\line(0,-1){1}}\put(6, 3){\line(-1,0){1}} \put(0,0){\line(1,0){2}}\put(2,0){\line(0,1){1}} \put(2,1){\line(1,0){1}} \multiput(3,1)(0.2,0.2){10}{\mbox{$\cdot$}} \put(2.5,2.7){\framebox(0.75,0.75){$ j$}} \put(0, 3.05){\line(1,0){2.5}}\put(2.9, 5){\line(0,-1){1.55}} \put(-0.55, 2.9){\mbox{-b}}\put(2.9, 5.25){\mbox{a}} \put(5,1.5){\mbox{$j=a-b+i$}} \end{picture}\par \caption{Residue for Young diagram $|\lambda, i\rangle$} \end{figure} Let $K=\mathbb Q(r, s)$ be the field of rational functions in $r$ and $s$. For each $i\in \mathbb Z_n$ we define the Fock space $\mathcal F_i$ to be \begin{equation} \mathcal F_i=\bigoplus_{\lambda\in \mathcal P} K|\lambda, i\rangle. \end{equation} Let $\lambda$ and $\mu$ be two partitions such that $\mu$ is obtained from $\lambda$ by adding a node $\gamma$ of residue $i$, or $\mu/\lambda$ consists of a single $i$-node $\gamma$. Let $I_i(\lambda)$ be the set of indent $i$-nodes in the boundary of $\lambda$, and $R_i(\lambda)$ be the set of its removable $i$-nodes in the boundary of $\lambda$. We further let $I_i^{(l)}(\lambda,\mu)$ (resp. $R_i^{(l)}(\lambda,\mu)$) be the set of indent $i$-nodes (resp. of removable $i$-nodes) situated to the left of $\gamma$ ($\gamma$ not included) in the boundary of $\mu$, and similarly, let $I_i^{(r)}(\lambda,\mu)$ and $R_i^{(r)}(\lambda,\mu)$ be the corresponding set of $i$-nodes located not to the left of $\gamma$ (i.e. to the right of $\gamma$ and $\gamma$ itself) on the boundary of $\lambda$. We also denote by $I^{(0)}(\lambda)$ ( resp. $R^{(0)}(\lambda)$ ) the total number of indent (resp. removable) $0$-nodes in diagram $\lambda$. We denote by $|I_i^{(r)}(\lambda,\mu)|$ and $|R_i^{(r)}(\lambda,\mu)|$ the number of the set $I_i^{(r)}(\lambda,\mu)$ and $R_i^{(r)}(\lambda,\mu)$ respectively.\\[6pt] For a finite set $S$, we use $|S|$ to denote its cardinality. We define the action of the simple generators as follows. \begin{eqnarray*} && f_i~|~\lambda~ \rangle = \sum_{\mu} r^{|I_i^{(r)}(\lambda,~\mu)|}~s^{|R_i^{(r)}(\lambda,~\mu)|}~| ~\mu ~\rangle ,\\ && e_i~|~\lambda~ \rangle = \sum_{\mu} r^{|R_i^{(l)}(\mu,~ \lambda)|}~s^{|I_i^{(l)}(\mu, ~\lambda)|}~| ~\mu ~\rangle ,\\ && \omega_i~|~\lambda~ \rangle = r^{|I_i(\lambda)|}~s^{|R_i(\lambda)|}~| ~\lambda ~\rangle ,\\[6pt] &&\omega'_i~|~\lambda~ \rangle = r^{|R_i(\lambda)|}~s^{|I_i(\lambda)|}~| ~\lambda ~\rangle,\\[6pt] && D~|~\lambda~ \rangle = r^{- |I^0(\lambda)|}~s^{-|R^0(\lambda)|}~| ~\lambda ~\rangle,\\[6pt] && D'_i~|~\lambda~ \rangle = r^{-|R^0(\lambda)|}~s^{-|I^0(\lambda)|}~| ~\lambda ~\rangle, \end{eqnarray*} where the first sum runs through all $\mu$ such that $\mu/\lambda$ is a $i$-node, and the second sum runs through all $\mu$ such that $\lambda/\mu$ is a $i$-node. We remark that all the sums are obviously finite since it runs through the cells of the Young diagram. We also note that when $r=s^{-1}=q$, the above action reduces to that of the quantum affine algebra $U_q(\widehat{sl}_n)$ formulated in \cite{LT}. \begin{theorem} For each $i\in\mathbb Z_n$ the Fock space $\mathcal F_i$ is a level one representation of $\mathcal {U}_{r,s}(\hat{sl}_n) $ and contains $V(\Lambda_i)$ as a submodule and the highest weight vector is $|\phi, i>$, where $\phi$ is the empty diagram. \end{theorem} \noindent{\bf Proof.} First of all we note the statement about the submodule is obtained by computing the action of the Heisenberg subalgebra. By symmetry it is enough to show the statement for the basic module $V(\Lambda_0)$, so we will drop the second index $i=0$ in $|\lambda, 0\rangle$. Namely, we take $i=0$ and the residue for the node $(a, -b)$ is $a-b$. (i) From the construction we first have \begin{eqnarray*} \omega_j e_i~|~\lambda~\rangle &= &\omega_j \sum_{\gamma \in R_i(\lambda)}r^{|R_i^{(l)}(\lambda-\gamma,~\lambda)|} ~s^{|I_i^{(l)}(\lambda-\gamma,~\lambda)|} ~|~\lambda-\gamma~\rangle\\ &=& \sum_{\gamma \in R_i(\lambda)}r^{|R_i^{(l)}(\lambda-\gamma,~\lambda)|} ~s^{|I_i^{(l)}(\lambda-\gamma,~\lambda)|}r^{|I_j(\lambda-\gamma)|}~s^{|R_j(\lambda-\gamma)|}~ |~\lambda-\gamma~\rangle, \end{eqnarray*} and similarly \begin{equation*} e_i \omega_j~|~\lambda~\rangle=\sum_{\gamma \in R_i(\lambda)}r^{|R_i^{(l)}(\lambda-\gamma,~\lambda)|} ~s^{|I_i^{(l)}(\lambda-\gamma,~\lambda)|} ~ r^{|I_j(\lambda)|} ~s^{|R_j(\lambda)|}~|~\lambda-\gamma~\rangle. \end{equation*} Since for $\gamma\in R_i(\lambda)$, \begin{eqnarray} \label{1} |I_i(\lambda-\gamma)|= |I_i(\lambda)|+1,~~~|R_i(\lambda-\gamma)| =|R_i(\lambda)|-1, \end{eqnarray} It follows immediately from (\ref{1}) that $$\omega_i e_i = rs^{-1} e_i \omega_i = \langle i,\,i\rangle e_i \omega_i.$$ On the other hand it follows from $\gamma \in R_i(\lambda)~~~(i\neq j) $ that, \begin{eqnarray} &&|I_j(\lambda-\gamma)|= |I_j(\lambda)|+ \begin{cases}(\varepsilon_j,\, \alpha_i), \quad\ \; (i\in I_0,\, j\in I)\vspace{6pt} \cr {-(\varepsilon_{i+1},\, \alpha_0)}, \quad (i\in I_0,\,j=0)\end{cases}\\[6pt] &&|R_j(\lambda-\gamma)| =|R_j(\lambda)|+\begin{cases}{(\varepsilon_{j+1},\, \alpha_i)}, \quad\ \; (i\in I_0,\, j\in I) \vspace{6pt} \cr {(\varepsilon_1,\, \alpha_i)}, \quad (i\in I_0,\,j=0).\end{cases} \end{eqnarray} These combinatorial identities imply that $$\omega_j e_i = <i,~j> e_i \omega_j.$$ (ii) To check the commutation relation (A4), we consider the actions of $e_i$ and $f_i$: \begin{eqnarray*} && e_i f_i ~|~\lambda~\rangle = e_i \sum_{\gamma \in I_i(\lambda)}r^{|I_i^{(r)}(\lambda,~\lambda+\gamma)|} ~s^{|R_i^{(r)}(\lambda,~\lambda+\gamma)|} ~|~\lambda+\gamma~\rangle\\ && = \sum_{ {\gamma \in I_i(\lambda),}\atop {\gamma' \in R_i(\lambda+\gamma)}}\begin{array}{l}r^{|R_i^{(r)}(\lambda,~\lambda+\gamma)|} ~s^{|R_i^{(r)}(\lambda,~\lambda+\gamma)|}\\ \times~ r^{|R_i^{(l)}(\lambda+\gamma-\gamma',\lambda+\gamma)|} ~s^{|I_i^{(l)}(\lambda+\gamma-\gamma',\lambda+\gamma)|} |\lambda+\gamma-\gamma'\rangle \end{array}\\ && = \sum_{ {\gamma \in I_i(\lambda),}\atop {\gamma' \in R_i(\lambda)\cup I_i(\lambda)}} \begin{array}{l} r^{|I_i^{(r)}(\lambda,~\lambda+\gamma)|+|R_i^{(l)}(\lambda+\gamma-\gamma',\lambda+\gamma)|} \\ ~\times s^{|R_i^{(r)}(\lambda,~\lambda+\gamma)|+|I_i^{(l)}(\lambda+\gamma-\gamma',\lambda+\gamma)|}~ ~|~\lambda+\gamma-\gamma'\rangle \end{array}\\ && = \sum_{\gamma \in I_i(\lambda)}r^{|I_i^{(r)}(\lambda, \lambda+\gamma)|+|R_i^{(l)}(\lambda,\lambda+\gamma)|~ s^{|R_i^{(r)}}(\lambda,\lambda+\gamma)|+|I_i^{(l)}(\lambda,\lambda+\gamma)|} ~~|~\lambda\rangle \\ &&~~~~~ +\sum_{ {\gamma \in I_i(\lambda),}\atop {\gamma' \in R_i(\lambda)}} \begin{array}{l}r^{|I_i^{(r)}(\lambda,~\lambda+\gamma)|+|R_i^{(l)}(\lambda+\gamma-\gamma',\lambda+\gamma)|} \\~ \times s^{|R_i^{(r)}(\lambda,~\lambda+\gamma)|+|I_i^{(l)}(\lambda+\gamma-\gamma',\lambda+\gamma)|}~ |\lambda+\gamma-\gamma'\rangle\end{array}, \end{eqnarray*} where we used the result: $R_i(\lambda+\gamma)=R_i(\lambda)+\{ \gamma \}$. Reversing the order of the product we have, \begin{eqnarray*} && f_i e_i ~|~\lambda~\rangle = \sum_{ {\gamma' \in R_i(\lambda),}\atop {\gamma \in I_i(\lambda-\gamma')}} \begin{array}{l} r^{|R_i^{(l)}(\lambda-\gamma', ~\lambda)|} ~s^{|I_i^{(r)}(\lambda-\gamma',~\lambda)|}\\ ~\times~ r^{|I_i^{(r)}(\lambda-\gamma',\lambda-\gamma'+\gamma)|} ~s^{|R_i^{(r)}(\lambda-\gamma',\lambda-\gamma'+\gamma)|}~ |~\lambda-\gamma'+\gamma~\rangle \end{array}\\ && = \sum_{\gamma' \in R_i(\lambda)}r^{|R_i^{(l)}(\lambda-\gamma', \lambda)|+|I_i^{(r)}(\lambda-\gamma',\lambda)|~ s^{|I_i^{(l)}}(\lambda-\gamma',\lambda)|+|R_i^{(r)}(\lambda-\gamma',\lambda)|} ~|~\lambda~\rangle \\ &&~~~~~~~ +\sum_{ {\gamma' \in R_i(\lambda),} \atop {\gamma' \in I_i(\lambda)}} \begin{array}{l} r^{|R_i^{(l)}(\lambda-\gamma',~\lambda)|+|I_i^{(r)}(\lambda-\gamma',\lambda-\gamma'+\gamma)|} \\ ~\times~s^{|I_i^{(r\l)}(\lambda-\gamma',~\lambda)|+|R_i^{(r)}(\lambda-\gamma',\lambda-\gamma'+\gamma)|}~ |~\lambda-\gamma'+\gamma~\rangle\end{array},\\ \end{eqnarray*} The following fact is easily verified. \noindent{\bf Claim A}\, For all $\gamma \in I_i(\lambda),~~ \gamma' \in R_i(\lambda)$ \\ \begin{eqnarray} {\label{2}} && I_i^{(r)}(\lambda,\lambda+\gamma) -I_i^{(r)}(\lambda-\gamma', \lambda-\gamma'+\gamma) \\ &&\hskip2cm = R_i^{(l)}(\lambda-\gamma')- R_i^{(l)}(\lambda +\gamma-\gamma',\lambda+\gamma);\nonumber\\ {\label{3}} && I_i^{(l)}(\lambda-\gamma',\lambda)-I_i^{(l)}(\lambda+\gamma-\gamma',\lambda+ \gamma)\\ &&\hskip2cm = R_i^{(r)}(\lambda,\lambda+\gamma)-R_i^{(r)}(\lambda-\gamma',\lambda-\gamma'+\gamma).\nonumber \end{eqnarray} Combining the above two expressions in Claim A, we get that \begin{eqnarray*} & & [\,e_i,\, f_i\,] ~|~\lambda~\rangle \\ & =& \sum_{\gamma \in I_i(\lambda)}r^{|I_i^{(r)}(\lambda,\,~\lambda+\gamma)|+|R_i^{(l)}(\lambda,\,~\lambda+\gamma)|} ~s^{|I_i^{(l)}(\lambda,\,~\lambda+\gamma)|+|R_i^{(r)}(\lambda,\,~\lambda+\gamma)|}|~\lambda~\rangle\\ && - \sum_{\gamma' \in R_i(\lambda)}r^{|I_i^{(r)}(\lambda-\gamma',\,~\lambda)|+|R_i^{(l)}(\lambda-\gamma',\,~\lambda)|} ~s^{|I_i^{(l)}(\lambda-\gamma',\,~\lambda)|+|R_i^{(r)}(\lambda-\gamma',\,~\lambda)|}|~\lambda~\rangle \end{eqnarray*} The following Claim B is important for the further deduction.\\ \noindent{\bf Claim B} \,\, For all $\gamma \in I_i(\lambda),~~ \gamma' \in R_i(\lambda)$ \\ \begin{eqnarray} && |I_i(\lambda-\gamma')|=|I_i^{(r)}(\lambda-\gamma',\,\lambda)| +|I_i^{(l)}(\lambda-\gamma', \lambda)|+1;\\[3pt] && |R_i(\lambda+\gamma)|=|R_i^{(r)}(\lambda,\,\lambda+\gamma)| +|R_i^{(l)}(\lambda,\,\lambda+\gamma)|+1;\\[3pt] && |I_i(\lambda)|=|I_i^{(r)}(\lambda,\,\lambda+\gamma)| +|I_i^{(l)}(\lambda,\,\lambda+\gamma)|+1;\\[3pt] && |R_i(\lambda)|=|R_i^{(r)}(\lambda-\gamma',\,\lambda)| +|R_i^{(l)}(\lambda-\gamma',\,\lambda)|+1; \end{eqnarray} \medskip We note that the coefficient is given by \begin{eqnarray*} && \sum_{\gamma \in I_i(\lambda)}r^{|I_i^{(r)}(\lambda,\,~\lambda+\gamma)|+|R_i^{(l)}(\lambda,\,~\lambda+\gamma)|} ~s^{|I_i^{(l)}(\lambda,\,~\lambda+\gamma)|+|R_i^{(r)}(\lambda,\,~\lambda+\gamma)|}\\[3pt] &-& \sum_{\gamma' \in R_i(\lambda)}r^{|I_i^{(r)}(\lambda-\gamma',\,~\lambda)|+|R_i^{(l)}(\lambda-\gamma',\,~\lambda)|} ~s^{|I_i^{(l)}(\lambda-\gamma',\,~\lambda)|+|R_i^{(r)}(\lambda-\gamma',\,~\lambda)|}\\[3pt] &=&\sum_{\gamma \in I_i(\lambda)}r^{|I_i^{(r)}(\lambda,\,~\lambda+\gamma)|-|R_i^{(r)}(\lambda,\,~\lambda+\gamma)|+|R_i(\lambda+\gamma)|-1} ~s^{|I_i(\lambda)|-|I_i^{(r)}(\lambda,\,~\lambda+\gamma)|+|R_i^{(r)}(\lambda,\,~\lambda+\gamma)|-1}\\ &-& \sum_{\gamma' \in R_i(\lambda)}r^{|I_i^{(r)}(\lambda-\gamma',\,~\lambda)|-|R_i^{(r)}(\lambda-\gamma',\,~\lambda)|+|R_i(\lambda)|-1} ~s^{|I_i(\lambda-\gamma')|-|I_i^{(r)}(\lambda-\gamma',\,~\lambda)|+|R_i^{(r)}(\lambda-\gamma',\,~\lambda)|-1} \end{eqnarray*} where we used the relations (3.8) and (3.9) in the first term and the relations (3.7) and (3.10) in the second term. Then we have \begin{eqnarray*} \text{The action of}\quad [e_i, \,f_i]&=&\sum_{\gamma \in I_i(\lambda)}r^{|R_i(\lambda+\gamma)|}s^{|I_i(\lambda)|-2} (rs^{-1})^{|I_i^{(r)}(\lambda,\,~\lambda+\gamma)|-|R_i^{(r)}(\lambda,\,~\lambda+\gamma)|-1}\\[3pt] &-& \sum_{\gamma' \in R_i(\lambda)}r^{|R_i(\lambda)|+1}s^{|I_i(\lambda-\gamma')|-3} (rs^{-1})^{|I_i^{(r)}(\lambda-\gamma',\,\lambda)|-|R_i^{(r)}(\lambda-\gamma',\,\lambda)|-2}\\[3pt] &=&r^{|R_i(\lambda)|+1}s^{|I_i(\lambda)|-2}\Big(\sum_{\gamma \in I_i(\lambda)} (rs^{-1})^{|I_i^{(r)}(\lambda,\,~\lambda+\gamma)|-|R_i^{(r)}(\lambda,\,~\lambda+\gamma)|-1}\\[3pt] &&\hskip3cm - \sum_{\gamma' \in R_i(\lambda)} (rs^{-1})^{|I_i^{(r)}(\lambda-\gamma',\,\lambda)|-|R_i^{(r)}(\lambda-\gamma',\,\lambda)|-2}\Big)\\[3pt] &=&r^{|R_i(\lambda)|+1}s^{|I_i(\lambda)|-2}\Big( (rs^{-1})^{|I_i(\lambda)|-|R_i(\lambda)|-2}+(rs^{-1})^{|I_i(\lambda)|-|R_i(\lambda)|-3}\\[3pt] &&\hskip3.2cm +\cdots + (rs^{-1})^{|I_i(\lambda)|-|R_i(\lambda)|-|I_i(\lambda)|+|R_i(\lambda)|-1}\Big)\\[3pt] &=&\frac{r^{|I_i(\lambda)|}s^{|R_i(\lambda)|}-r^{|R_i(\lambda)|}s^{|I_i(\lambda)|}}{r-s} =\text{the action of} \quad \frac{\omega_i-\omega'_i}{r-s} \end{eqnarray*} Finally, we will check the $(r, s)-$ Serre relation: $$f_{i+1}f_i^2-(r+s)\,f_if_{i+1}f_i+rs\,f_i^2f_{i+1}=0.$$ It follows from the definition that \begin{eqnarray*} && f_{i+1}f_i^2~|~\lambda~\rangle \\ &=& \sum_{\mbox{\tiny $\begin{array}{c}\gamma_1 \in I_i(\lambda) \\ \gamma_2\in I_i(\lambda+\gamma_1) \\ \gamma_3\in I_{i+1}(\lambda +\gamma_1+\gamma_2)\end{array}$}} \begin{array}{l} r^{|I_i^{(r)}(\lambda,\, \lambda+\gamma_1)|+|I_i^{(r)}(\lambda+\gamma_1,\, \lambda+\gamma_1+\gamma_2)| +|I_{i+1}^{(r)}(\lambda +\gamma_1+\gamma_2,\, \lambda +\gamma_1+\gamma_2+\gamma_3)|}\\[3pt] s^{|R_i^{(r)}(\lambda,\, \lambda+\gamma_1)|+ |R_i^{(r)}(\lambda+\gamma_1,\, \lambda +\gamma_1+\gamma_2)|+ |R_{i+1}^{(r)}(\lambda+\gamma_1+\gamma_2,\, \lambda+\gamma_1+\gamma_2+\gamma_3)|}\\[3pt] ~~\,\,|~~\lambda+\gamma_1+\gamma_2+\gamma_3\rangle \end{array} \end{eqnarray*} For simplicity, we write $$I_{i+1}(\lambda +\gamma_1+\gamma_2)=I_{i+1}(\lambda)\cup \bigg(I_{i+1}(\lambda +\gamma_1)-I_{i+1}(\lambda)\bigg)\cup \bigg(I_{i+1}(\lambda +\gamma_2)-I_{i+1}(\lambda)\bigg),$$ which is derived since $I_{i+1}(\lambda+\gamma_1+\gamma_2)= I_{i+1}(\lambda)\cup I_{i+1}(\lambda+\gamma_1)\cup I_{i+1}(\lambda+\gamma_2)$. Then the coefficient of the above expression becomes \begin{eqnarray*} & & \sum_{\mbox{\tiny $\begin{array}{c}\gamma_1 \in I_i(\lambda)\\ \gamma_2\in I_i(\lambda+\gamma_1) \\ \gamma_3\in I_{i+1}(\lambda)\end{array}$}} \begin{array}{l}r^{|I_i^{(r)}(\lambda,\,\lambda+\gamma_1)|+|I_i^{(r)}(\lambda+\gamma_1,\, \lambda+\gamma_1+\gamma_2)| +|I_{i+1}^{(r)}(\lambda,\, \lambda+\gamma_3)|}\\[6pt] \times s^{|R_i^{(r)}(\lambda,\lambda+\gamma_1)|+ |R_i^{(r)}(\lambda+\gamma_1,\, \lambda +\gamma_1+\gamma_2)|+ |R_{i+1}^{(r)}(\lambda,\, \lambda+\gamma_3)|}\end{array}\\ && {\displaystyle \hskip 5mm + \sum_{\mbox{\tiny $\begin{array}{c}\gamma_1 \in I_i(\lambda)\\ \gamma_2\in I_i(\lambda+\gamma_1) \\ \gamma_3\in I_{i+1}(\lambda +\gamma_1) - I_{i+1}(\lambda)\end{array}$}} \begin{array}{l} r^{|I_i^{(r)}(\lambda,\, \lambda+\gamma_1)|+|I_i^{(r)}(\lambda+\gamma_1,\, \lambda+\gamma_1+\gamma_2)| +|I_{i+1}^{(r)}(\lambda +\gamma_1,\, \lambda +\gamma_1+\gamma_3)|+1}\\[6pt] \times s^{|R_i^{(r)}(\lambda,\, \lambda+\gamma_1)|+ |R_i^{(r)}(\lambda+\gamma_1,\, \lambda +\gamma_1+\gamma_2)|+ |R_{i+1}^{(r)}(\lambda+\gamma_1,\, \lambda+\gamma_1+\gamma_3)|}\end{array}}\\ && {\displaystyle \hskip 5mm + \sum_{\mbox{\tiny $\begin{array}{c}\gamma_1 \in I_i(\lambda)\\ \gamma_2\in I_i(\lambda+\gamma_1) \\ \gamma_3\in I_{i+1}(\lambda+\gamma_2)- I_{i+1}(\lambda)\end{array}$}} \begin{array}{l} r^{|I_i^{(r)}(\lambda,\, \lambda+\gamma_1)|+|I_i^{(r)}(\lambda+\gamma_1,\, \lambda+\gamma_1+\gamma_2)| +|I_{i+1}^{(r)}(\lambda+\gamma_2,\, \lambda+\gamma_2+\gamma_3)|+1}\\[6pt] \times s^{|R_i^{(r)}(\lambda,\, \lambda+\gamma_1)|+ |R_i^{(r)}(\lambda+\gamma_1,\, \lambda +\gamma_1+\gamma_2)|+ |R_{i+1}^{(r)}(\lambda+\gamma_2,\, \lambda+\gamma_2+\gamma_3)|}\end{array}}\\ \end{eqnarray*} Furthermore we get \begin{eqnarray*} & & f_i f_{i+1}f_i~|~\lambda~\rangle \\ && = \sum_{\mbox{\tiny $\begin{array}{c}\gamma_1 \in I_i(\lambda)\\ \gamma_3\in I_{i+1}(\lambda+\gamma_1) \\ \gamma_2\in I_i(\lambda +\gamma_1+\gamma_3)\end{array}$}} \begin{array}{l} r^{|I_i^{(r)}(\lambda,\, \lambda+\gamma_1)|+|I_{i+1}^{(r)}(\lambda+\gamma_1,\, \lambda+\gamma_1+\gamma_3)| +|I_i^{(r)}(\lambda +\gamma_1+\gamma_3,\, \lambda +\gamma_1+\gamma_3+\gamma_2)|}\\[3pt] s^{|R_i^{(r)}(\lambda,\, \lambda+\gamma_1)|+ |R_{i+1}^{(r)}(\lambda+\gamma_1,\, \lambda +\gamma_1+\gamma_3)|+ |R_i^{(r)}(\lambda+\gamma_1+\gamma_3,\, \lambda+\gamma_1+\gamma_3+\gamma_2)|}\\[3pt]~~\,\,|~~\lambda+\gamma_1+\gamma_2+\gamma_3\rangle\end{array} \end{eqnarray*} Similarly, using $$I_{i+1}(\lambda +\gamma_1)=I_{i+1}(\lambda)\cup \bigg(I_{i+1}(\lambda +\gamma_1)-I_{i+1}(\lambda)\bigg),$$ the coefficient of the second expression becomes \begin{eqnarray*} & & \sum_{\mbox{\tiny $\begin{array}{c}\gamma_1 \in I_i(\lambda)\\ \gamma_3\in I_{i+1}(\lambda) \\ \gamma_2\in I_i(\lambda+\gamma_1)\end{array}$}} \begin{array}{l}r^{|I_i^{(r)}(\lambda,\, \lambda+\gamma_1)|+|I_{i+1}^{(r)}(\lambda,\, \lambda+\gamma_3)| +|I_i^{(r)}(\lambda+\gamma_1,\, \lambda+\gamma_1+\gamma_2)|}\\[6pt] s^{|R_i^{(r)}(\lambda,\, \lambda+\gamma_1)|+ |R_{i+1}^{(r)}(\lambda,\, \lambda+\gamma_3)|+ |R_i^{(r)}(\lambda+\gamma_1,\, \lambda +\gamma_1+\gamma_2)|-1}\end{array}\\ && {\displaystyle \hskip 5mm + \sum_{\mbox{\tiny $\begin{array}{c}\gamma_1 \in I_i(\lambda)\\ \gamma_3\in I_{i+1}(\lambda+\gamma_1)-I_{i+1}(\lambda) \\ \gamma_2\in I_i(\lambda +\gamma_1)\end{array}$}} \begin{array}{l} r^{|I_i^{(r)}(\lambda,\, \lambda+\gamma_1)|+|I_{i+1}^{(r)}(\lambda+\gamma_1,\, \lambda+\gamma_1+\gamma_3)| +|I_i^{(r)}(\lambda +\gamma_1,\, \lambda +\gamma_1+\gamma_2)|+1}\\[6pt] s^{|R_i^{(r)}(\lambda,\, \lambda+\gamma_1)|+ |R_{i+1}^{(r)}(\lambda+\gamma_1,\, \lambda +\gamma_1+\gamma_3)|+ |R_i^{(r)}(\lambda+\gamma_1,\, \lambda+\gamma_1+\gamma_2)|-1}\end{array}} \end{eqnarray*} Finally, using the definition again one has \begin{eqnarray*} & & f_i^2 f_{i+1}~|~\lambda~\rangle \\ && = \sum_{\mbox{\tiny $\begin{array}{c}\gamma_3 \in I_{i+1}(\lambda)\\ \gamma_1\in I_i(\lambda+\gamma_3) \\ \gamma_2\in I_i(\lambda +\gamma_1+\gamma_3)\end{array}$}} \begin{array}{l} r^{|I_{i+1}^{(r)}(\lambda,\, \lambda+\gamma_3)|+|I_i^{(r)}(\lambda+\gamma_3,\, \lambda+\gamma_3+\gamma_1)| +|I_i^{(r)}(\lambda +\gamma_1+\gamma_3,\, \lambda +\gamma_1+\gamma_3+\gamma_2)|}\\[3pt] s^{|R_{i+1}^{(r)}(\lambda,\, \lambda+\gamma_1)|+ |R_i^{(r)}(\lambda+\gamma_3,\, \lambda +\gamma_3+\gamma_1)|+ |R_i^{(r)}(\lambda+\gamma_1+\gamma_3,\, \lambda+\gamma_1+\gamma_2+\gamma_3)|}\\[3pt]~~\,\,|~~\lambda+\gamma_1+\gamma_2+\gamma_3\rangle\end{array} \end{eqnarray*} Since for $\gamma_3\in I_{i+1}(\lambda)$, one gets $$I_i(\lambda+\gamma_3)=I_i(\lambda), \qquad I_i(\lambda +\gamma_1+\gamma_3)=I_i(\lambda +\gamma_1).$$ Thus the coefficient of the third expression becomes \begin{eqnarray*} && \sum_{\mbox{\tiny $\begin{array}{c}\gamma_3 \in I_{i+1}(\lambda)\\ \gamma_1\in I_i(\lambda) \\ \gamma_2\in I_i(\lambda +\gamma_1)\end{array}$}} \begin{array}{l} r^{|I_{i+1}^{(r)}(\lambda,\, \lambda+\gamma_3)|+|I_i^{(r)}(\lambda,\, \lambda+\gamma_1)| +|I_i^{(r)}(\lambda +\gamma_1,\, \lambda +\gamma_1+\gamma_2)|}\\[3pt] \times s^{|R_{i+1}^{(r)}(\lambda,\, \lambda+\gamma_3)|+ |R_i^{(r)}(\lambda,\, \lambda +\gamma_1)|-1+ |R_i^{(r)}(\lambda+\gamma_1,\, \lambda+\gamma_1+\gamma_2)|-1}\end{array}\\ \end{eqnarray*} Combining the above three coefficients, we get the required result: $$f_{i+1}f_i^2-(r+s)\,f_if_{i+1}f_i+rs\,f_i^2f_{i+1}|\lambda \rangle=0,$$ where we have used the following fact.\\[6pt] \noindent{\bf Claim C}\, \, For $\gamma_2\in I_i(\lambda+\gamma_1)$ and $\gamma_3\in I_{i+1}(\lambda+\gamma_1+\gamma_2)$, it follows that, \begin{eqnarray*} && | I_{i+1}^{(r)} (\lambda+\gamma_2, ~ \lambda +\gamma_2+\gamma_3)| = |I_{i+1}^{(r)} (\lambda+\gamma_1, ~ \lambda +\gamma_1+\gamma_3)|+1,\\[6pt]&& |R_{i+1}^{(r)} (\lambda+\gamma_2, ~ \lambda +\gamma_2+\gamma_3)| = |R_{i+1}^{(r)} (\lambda+\gamma_1, ~ \lambda +\gamma_1+\gamma_3)|-1. \end{eqnarray*} \medskip This completes the proof of Theorem 3.1. \vskip30pt \centerline{\bf ACKNOWLEDGMENT} \bigskip N. Jing would like to thank the support of NSA grant and NSFC (No. 10728102). H. Zhang would like to thank the support of NSFC (No. 10801094)and Shanghai Leading Academic Discipline Project(No. J50101). \bigskip \bibliographystyle{amsalpha}
\section{Introduction}\label{intro-scn} The analysis of shape may be counted among the very early activities of mankind; be it for representation on cultural artefacts, or for morphological, biological and medical applications. In modern days shape analysis is gaining increased momentum in computer vision, image analysis, biomedicine and many other fields. For a recent overview cf. \cite{KY06}. A \emph{shape space} can be viewed as the quotient of a Riemannian manifold -- e.g. the pre-shape sphere of centered unit size landmark configurations -- modulo the isometric and proper action of a Lie group (cf. \cite{Bre72}), conveying shape equivalence -- e.g. the group of rotations, cf. \citep[Chapter 11]{KBCL99}. Thus, it carries the canonical quotient structure of a union of manifold strata of different dimensions, which give in general a Riemannian \emph{manifold part} -- possibly with singularities comprising the \emph{non-manifold part} of \emph{non-regular shapes} at some of which sectional curvatures may tend to infinity, cf. \citep[Chapter 7.3]{KBCL99} as well as \cite{HHM07}. In a Euclidean space, there is a clear and unique concept of a mean in terms of least squares minimization: the arithmetic average. Generalizing to manifolds, however, the concept of expectation, average or \emph{mean} is surprisingly non trivial and not at all canonical. In fact, it resulted in an overwhelming number of different concepts of means, each defined by a specific concept of a distance, all of which are identical for the Euclidean distance in a Euclidean space. More precisely, with every embedding in a Euclidean space come specific \emph{extrinsic} and \emph{residual means} and with every Riemannian structure comes a specific \emph{intrinsic mean}. Furthermore, due to the non-Euclidean geometry, local minimizers introduced as \emph{Karcher means} by \cite{KWS90} may be different from global minimizers called \emph{Fr\'echet means} by \cite{Z77}, and, neither ones are necessarily unique. Nonetheless, carrying statistics over to manifolds, strong consistency (by \cite{Z77}, \cite{BP03}) and under suitable conditions, central limit theorems (CLTs) for such means have been derived (by \cite{J88}, \cite{HL96,HL98}, \cite{BP05} as well as \cite{H_Procrustes_10}). On shape spaces, various other concepts of means have been introduced, e.g. the famous \emph{Procrustes means} (cf. \cite{Z77},\cite{DM98}). As we show here, these means are related to the above ones via a \emph{horizontal lifting} from the bottom quotient to the top manifold, cf. Table \ref{means:tab}. In particular, since there are many -- and often confusing -- variants of Procrustes means in the literatur this paper introduces the terminology of \emph{Procrustean means} standing for inheritance from residual means. \begin{table}[h!] \centering \fbox{$\begin{array}{c|c} \mbox{manifold means} & \mbox{shape means}\\ \hline \mbox{intrinsic}&\mbox{intrinsic}\\ \mbox{extrinsic}&\mbox{Ziezold}\\ \mbox{residual} & \mbox{Procrustean} \end{array}$} \caption{\it Three fundamental types of means on a shape space (right column) and their horizontal lifts to the respective manifold (left column). \label{means:tab}} \end{table} For a CLT to hold, a manifold structure locally relating to a Euclidean space is sufficient. This leads to the question under which conditions it can be guaranteed that a mean shape lies on the manifold part. Due to strong consistency, for a \emph{one-sample test} for a specific mean shape on the manifold part, it may be assumed that sample means eventually lie on the manifold part as well, thus making the above cited CLTs available. To date however, a two- and a multi-sample test could not be justified because of a lacking result on the following manifold stability. \begin{Def} A mean shape enjoys \emph{manifold stability} if it is assumed on the manifold part for any random shape assuming the manifold part with non-zero probability. \end{Def} A key result of this paper establishes manifold stability for intrinsic and Ziezold means under the following condition. \begin{Cond}\label{realistic:cond} On the non-manifold part the distribution of the random shape contains at most countably many point masses. \end{Cond} Since the non-manifold part is a null-set (e.g. \cite{Bre72}) under the projection of the Riemannian volume, this condition covers most realistic cases. We develop the corresponding theory for a general shape space quotient based on lifting a distribution on the shape space to the pre-shape space and subsequently exploiting the fact that intrinsic means are zeroes of an integral involving the Riemann exponential. The similar argument can be applied to Ziezold means but not to Procrustean means. More specifically, we develop the notion of a \emph{measurable horizontal lift} of the shape space except for its \emph{quotient cut locus} (introduced as well) to the pre-shape space. This requires the geometric concept of \emph{tubular neighborhoods admitting slices}. Curiously, the result applied to the finite dimensional subspaces exhausting the quotient shape space of closed planar curves with arbitrary initial point introduced by \cite{ZR72} and further studied by \cite{KSMJ04}, gives that the shape of the circle, since it is a singularity, can never be an intrinsic shape mean of non-circular curves. As a second curiosity, 3D full Procrustes means do not enjoy manifold stability in general, a counterexample involving low concentration is given. This is due to the fact that for low concentration, full Procrustes means may be `blinder' in comparison to intrinsic and Ziezold means to distributional changes far away from a mode. Included in this context is also a discussion of the \emph{Schoenberg means}, recently introduced by \cite{BandPat05} as well as by \cite{DKLW08} for the non-manifold Kendall reflection shape spaces, which in the ambient space, also allow for a CLT. Schoenberg means, as demonstrated, however, may feature `blindness' in comparison to intrinsic and Ziezold means, with respect to changes in the distribution of nearly degenerate shapes. In a simulation we show that these features render Schoenberg means less effective for a discrimination involving degenerate or nearly degenerate shapes. As a third curiosity, for spheres and Kendall's shape spaces, it is shown that, given uniqueness, with order of concentration, the (generalized) geodesic segment between the intrinsic mean and the residual/Procrustean mean is in approximation divided by the extrinsic/Ziezold mean by the ratio $1:3$. This first order relationship can be readily observed in existing data sets. In particular, this result supports the conjecture that Procrustean means of sufficiently concentrated distributions enjoy stability as well. This paper is structured as follows. For convenience of the reader, first in Section \ref{Kendalls-ss:scn}, Kendall's shape spaces are introduced along with the specific result on manifold stability, followed by a classification of concepts of means on general shape spaces in Section \ref{Frechet-rho-means:scn}. The rather technical Section \ref{convex:scn} develops horizontal lifting and establishes manifold stability, technical proofs are deferred to the appendix. In Section \ref{ext-means:scn} extrinsic Schoenberg means are discussed and Section \ref{local:scn} tackles local effects of curvature on spheres and Kendall's shape spaces. Section \ref{class_data_simulations_scn} illustrates practical consequences using classical data-sets as well as simulations. Note that lacking stability does not affect the validity of the Strong Law, on which the considerations on asymptotic distance in Sections \ref{local:scn} and \ref{class_data_simulations_scn} are based. An R-package for all of the computations performed is provided online: \cite{Hshapes}. \section{Stability of Means on Kendall's Shape Spaces}\label{Kendalls-ss:scn} In the statistical analysis of similarity shapes based on landmark configurations, geometrical $m$-dimensional objects (usually $m=2,3$) are studied by placing $k>m$ \emph{landmarks} at specific locations of each object. Each object is then described by a matrix in the space $M(m,k)$ of $m\times k$ matrices, each of the $k$ columns denoting an $m$-dimensional landmark vector. $\langle x,y\rangle := \operatorname {tr}(xy^T)$ denotes the usual inner product with norm $\|x\| = \sqrt{\langle x,x\rangle}$. For convenience and without loss of generality for the considerations below, only \emph{centered} configurations are considered. Centering can be achieved by multiplying with a sub-Helmert matrix ${\cal H}\in M(k,k-1)$ from the right, yielding a configuration $x{\cal H}$ in $M(m,k-1)$. For this and other centering methods cf. \citep[Chapter 2]{DM98}. Excluding also all configurations with all landmarks coinciding gives the space of \emph{configurations} \begin{eqnarray* F_m^k&:=& M(m,k-1) \setminus \{0\} \,. \end{eqnarray*} Since only the similarity shape is of concern, we may assume that all configurations are contained in the unit sphere $ S_m^k :=\{x\in F_m^k: \|x\|=1\}$ called the \emph{pre-shape sphere}. Then, \emph{Kendall's shape space} is the canonical quotient $$\Sigma_m^k := S_m^k/SO(m) = \{[x]:x\in S_m^k\}\mbox{ with the \emph{orbit} } [x] = \{gx:g\in SO(m)\}\,.$$ In some applications reflections are also filtered out giving \emph{Kendall's reflection shape space} $$R\Sigma_m^k := \Sigma_m^k/\{e,\widetilde{e}\}= S_m^k/O(m)\,.$$ Here, $O(m) = \{g\in M(m,m): g^Tg=e\}$ denotes the orthogonal group with the unit matrix $e=\operatorname {diag}(1,\ldots,1)$, $\widetilde{e} = \operatorname {diag}(-1,1,\ldots,1)$ and $SO(m) = \{g\in O(m): \det(g)=1\} $ is the special orthogonal group. For $1\leq j < m < k$ consider the isometric embedding \begin{eqnarray}\label{pre-shape-emb:eq} \begin{array}{rclcrcl} S_j^k &\hookrightarrow& S_m^k&:& x&\mapsto&\left(\begin{array}{c}x\\\hline 0\end{array}\right)\end{array}\, \end{eqnarray} giving rise to a canonical embedding $R\Sigma_j^k\hookrightarrow \Sigma_m^k$ which is isometric w.r.t. the canonical intrinsic distance, the Procrustean distance and the Ziezold distance, respectively, defined in Section \ref{Frechet-rho-means:scn}, cf. \citep[p. 29]{KBCL99}, cf. also Remark \ref{sharp:rm} below. We say that a configuration in $\mathbb R^m$ is \emph{$j$-dimensional}, or more precisely \emph{non-degenerate $j$-dimensional} if its preshape $x\in S_m^k$ is of rank $j$. Moreover, for $j\geq 3$ the shape spaces $\Sigma_j^k$ and $R\Sigma_j^k$ decompose into a \emph{manifold part} (defined in Section \ref{Frechet-rho-means:scn}, cf. also Section \ref{ext-means:scn}) of \emph{regular shapes} $$(\Sigma_j^k)^* = \{[x]\in \Sigma_j^k: \operatorname {rank}(x) \geq j-1\}\mbox{ and }(R\Sigma_j^k)^* = \{[x]\in R\Sigma_j^k: \operatorname {rank}(x) =j\}\,,$$ respectively, given by the shapes corresponding to configurations of at least dimension $j-1$ and $j$, respectively and a non void part of singular shapes corresponding to lower dimensional configurations, respectively. The following Theorem for intrinsic means, full Procrustes means and Ziezold means (also defined in Section \ref{Frechet-rho-means:scn}) of random elements taking values in $R\Sigma_j^k$ follows from Proposition \ref{res_means_sphere:prop}, Remark \ref{kend-mf-quot-mean:rm} and the fact that $R\Sigma_j^k\subset \Sigma_m^k$ contains all shapes in $\Sigma_m^k$ of configurations of dimension up to $j$, $1\leq j<m<k$. \begin{Th}\label{decr_dim:th} Suppose that $X$ is a random shape on $\Sigma_m^k$ assumes shapes in $R\Sigma_j^k$ ($1\leq j < m<k$) with probability one. Then every full Procrustes mean shape of $X$ and every unique intrinsic or Ziezold mean shape under Condition \ref{realistic:cond} w.r.t. $(R\Sigma_j^k)^*$ corresponds to a configuration of dimension less than or equal to $j$. \end{Th} The following theorem is the application of the key result applied to Kendall's shape spaces. \begin{Th}[Stability Theorem for Intrinsic and Ziezold means \label{Kendall_mean_dim} Let $X$ be a random shape on $\Sigma_m^k$, $0<m<k$, with unique intrinsic or Ziezold mean shape $[\mu] \in \Sigma_m^k$, $\mu \in S_m^k$ and let $1\leq j \leq m$ be the maximal dimension of configurations of shapes assumed by $X$ with non-zero probability. Suppose moreover that shapes of configurations of strictly lower dimensions are assumed with at most countably many point masses. \begin{enumerate}\item[(i)] If $j<m$ then $\mu$ corresponds to a non-degenerate $j$-dimensional configuration. \item[(ii)]If $j=m$ then $\mu$ corresponds to a non-degenerate configuration of dimension $m-1$ or $m$. \end{enumerate} \end{Th} \begin{proof} Lemma \ref{cut_locus:lem} teaches that for Kendall's shape spaces, all quotient cut loci are void. Since for Ziezold means, Remark \ref{kendall_book:rm} provides invariant optimal positioning and Remark \ref{half-sphere:rm} provides the validity of (\ref{Ziez-inj:cond}), Corollary \ref{pop_mean_reg:cor} applied to $R\Sigma_{j}^k$ as well as to $\Sigma_m^k$ states that intrinsic and Ziezold means are also assumed on the manifold parts of $R\Sigma_{j}^k$ and $\Sigma_m^k$, respectively. In conjunction with Theorem \ref{decr_dim:th}, this gives the assertion. \end{proof} \begin{Rm}\label{sharp:rm} The result of Theorem \ref{Kendall_mean_dim} is sharp. To see this, consider for $\alpha >\beta >0$, $\alpha^2+\beta^2 =1$ the pre-shapes $$ x = \left(\begin{array}{ccc}\alpha &0&0\\0&\beta&0 \end{array}\right),~~y = \left(\begin{array}{ccc}\alpha &0&0\\0&-\beta&0 \end{array}\right)\mbox{ and }z = \left(\begin{array}{ccc}1 &0&0\\0&0&0 \end{array}\right)\, \in S_2^4\,.$$ Then $x$ and $y$ correspond to non-degenerate two-dimensional quadrilateral configurations while $z$ corresponds to a one-dimensional (collinear) quadrilateral. Still, $[z]$ is regular in $\Sigma_2^4$ and it is the intrinsic and Ziezold mean of $[x]$ and $[y]$ in $\Sigma_2^4$. Under the embedding $R\Sigma_2^4\hookrightarrow \Sigma_3^4$ we have the pre-shapes $$x' = \left(\begin{array}{ccc}\alpha &0&0\\0&\beta&0 \\0&0&0\end{array}\right),~~y' = \left(\begin{array}{ccc}\alpha &0&0\\0&-\beta&0\\0&0&0 \end{array}\right)\mbox{ and }z' = \left(\begin{array}{ccc}1 &0&0\\0&0&0 \\0&0&0\end{array}\right)\in S_3^4\,.$$ Just as $[x] = [y]$ in $R\Sigma_2^4$ so do $x'$ and $y'$ have regular and identical shape in $\Sigma_3^4$. However, $[z']$ is not regular and it is not the intrinsic or Ziezold mean in $\Sigma_3^4$. \end{Rm} \section{Fundamental Types of Means} \label{Frechet-rho-means:scn} In the previous section we introduced Kendall's \emph{shape} and \emph{reflection shape space} based on invariance under similarity transformations and, including reflections, respectively. Invariance under congruence transformations only leads to Kendall's \emph{size-and-shape space}. More generally in image analysis, invariance may also be considered under the affine or projective group, cf. \cite{MP01,MP05}. A different yet also very popular popular set of shape spaces for two-dimensional configurations modulo the group of similarities has been introduced by \cite{ZR72}. Instead of building on a finite dimensional Euclidean matrix space modeling landmarks, the basic ingredient of these spaces modeling closed planar unit speed curves is the infinite dimensional Hilbert space of Fourier series, cf. \cite{KSMJ04}. In practice for numerical computations, only finitely many Fourier coefficients are considered. To start with, a shape space is a metric space $(Q,d)$. For this entire paper suppose that $X, X_1,X_2,\ldots$ are i.i.d. random elements mapping from an abstract probability space $(\Omega,\cal A,\operatorname {\mathbb P})$ to $(Q,d)$ equipped with its self understood Borel $\sigma$-field. Here and in the following, \emph{measurable} will refer to the corresponding Borel $\sigma$-algebras, respectively. Moreover, denote by $\mathbb E(Y)$ the classical expected value of a random vector $Y$ on a $D$-dimensional Euclidean space $\mathbb R^D$, if existent. \begin{Def}\label{Frechet_means:def} For a continuous function $\rho:Q\times Q \to [0,\infty)$ define the \emph{set of population Fr\'echet $\rho$-means} by $$ E^{(\rho)}(X) = \operatorname {argmin}_{\mu\in Q} \mathbb E\big(\rho(X,\mu)^2\big) \,.$$ For $\omega\in \Omega$ denote the \emph{set of sample Fr\'echet $\rho$-means} by $$ E^{(\rho)}_n(\omega) = \operatorname {argmin}_{\mu\in Q} \sum_{j=1}^n \rho\big(X_j(\omega),\mu\big)^2\,.$$ \end{Def} By continuity of $\rho$, the $\rho$-means are closed sets, additionally, sample $\rho$-means are random sets, all of which may be empty. For our purpose here, we rely on the definition of \emph{random closed sets} as introduced and studied by \cite{Choq54}, \cite{Kend74} and \cite{Math75}. Since their original definition for $\rho=d$ by \cite{F48} such means have found much interest. \paragraph{Intrinsic means.} Independently, for a connected Riemannian manifold with geodesic distance $\rho^{(i)}$, \cite{KN69} defined the corresponding means as \emph{centers of gravity}. They are nowadays also well known as \emph{intrinsic means} by \cite{BP03,BP05}. \paragraph{Extrinsic means.} W.r.t. the chordal or \emph{extrinsic metric} $\rho^{(e)}$ due to an embedding of a Riemannian manifold in an ambient Euclidean space, Fr\'echet $\rho$-means have been called \emph{mean locations} by \cite{HL96} or \emph{extrinsic means} by \cite{BP03}. More precisely, let $Q=M\subset \mathbb R^D$ be a complete Riemannian manifold embedded in a Euclidean space $\mathbb R^D$ with standard inner product $\langle x,y\rangle$, $\|x\|=\sqrt{\langle x,x\rangle}, \rho^{(e)}(x,y)=\|x-y\|$ and let $\Phi: \mathbb R^D \to M$ denote the orthogonal projection, $\Phi(x) = {\rm argmin}_{p\in M}\|x-p\|$. For any Riemannian manifold an embedding that is even isometric can be found for $D$ sufficiently large, see \cite{Na56}. Due to an extension of Sard's Theorem by \citep[p.12]{BP03} for a closed manifold, $\Phi$ is univalent up to a set of Lebesgue measure zero. Then the set of extrinsic means is given by the set of images $\Phi\big(\mathbb E(Y)\big)$ where $Y$ denotes $X$ viewed as taking values in $\mathbb R^D$ (cf. \cite{BP03}). \paragraph{Residual means.} In this context, setting $\rho^{(r)}(p,{p'}) = \|d\Phi_{p'}(p-{p'})\|$ ($p,{p'}\in M$) with the derivative $d\Phi_{p'}$ at ${p'}$ yielding the orthogonal projection to the embedded tangent space $T_{p'}\mathbb R^D \to T_{p'}M\subset T_{p'}\mathbb R^D$, call the corresponding mean sets $E^{(\rho^{(r)})}(X)$ and $E^{(\rho^{(r)})}_n(\omega)$, the sets of \emph{residual population means} and \emph{residual sample means}, respectively. For two-spheres, $\rho^{(r)}(p,{p'})$ has been studied under the name of \emph{crude residuals} by \cite{J88}. On unit-spheres \begin{eqnarray}\label{res_dist_sphere:def} \rho^{(r)}(p,{p'}) =\|p - \langle p,{p'}\rangle {p'}\| = \sqrt{1-\langle p,{p'}\rangle^2} = \rho^{(r)}({p'},p) \end{eqnarray} is a quasi-metric (symmetric, vanishing on the diagonal $p={p'}$ and satisfying the triangle inequality). On general manifolds, however, the \emph{residual distance} $\rho^{(r)}$ may be neither symmetric nor satisfying the triangle inequality. ~\\ Obviously, for $X$ uniformly distributed on a sphere, the entire sphere is identical with the set of intrinsic, extrinsic and residual means: non-unique intrinsic and extrinsic means may depend counterintuitively on the dimension of the ambient space. Here is a simple illustration. \begin{Prop}\label{spherical_means:thm} Suppose that $X$ is a random point on a unit sphere $S^{D-1}$ that is uniformly distributed on a unit subsphere $S$. Then \begin{enumerate} \item[(i)] every point on $S^{D-1}$ is an extrinsic mean and, \item[(ii)] if $S$ is a proper subsphere then the set of intrinsic means is equal to the unit subsphere $S'$ orthogonal to $S$. \end{enumerate} \end{Prop} \begin{proof} The first assertion is a consequence of $ \rho^{(e)}(x,y)^2 + \rho^{(e)}(x,-y)^2 = 4$ for every $x,y\in S^{D-1}$. The second assertion follows from $$ \rho^{(i)}(x,y)^2 + \rho^{(i)}(x,-y)^2 = \rho^{(i)}(x,y)^2 + \big(\pi -\rho^{(i)}(x,y)\big)^2~\geq~ \frac{\pi^2}{2} \mbox{ for every }x,y\in S^{D-1}\,$$ for the intrinsic distance $\rho^{(i)}(x,y) = 2\arcsin(\|x-y\|/2)$ with equality if and only if $x$ is orthogonal to $y$. \end{proof} \begin{Prop}\label{res_means_sphere:prop} If a random point $X$ on a unit sphere is a.s. contained in a unit subsphere $S$ then $S$ contains every residual mean as well as every unique intrinsic or extrinsic mean .\end{Prop} \begin{proof} Suppose that $x = v+\nu$ is a mean of $X$ with $v/\|v\| \in S$ and $\nu\in S^{D-1}$ normal to $S$. Since $1-\langle X,v+\nu\rangle^2 = 1-\langle X,v\rangle^2 \geq 1 - \langle X,v\rangle^2/\|v\|^2$ a.s. with equality if and only if $\nu=0$, the assertion for residual means follows at once from (\ref{res_dist_sphere:def}). For intrinsic and extrinsic means we argue with $\|X-(v+\nu)\| = \|X-(v-\nu)\|$ a.s. yielding $\nu=0$ in case of uniqueness. \end{proof} Let us now incorporate more of the structure common to shape spaces. The following definition is due to \citep[p. 249]{KBCL99}. We additionally require that the group acting be compact in order to ensure that the quotient be Hausdorff. More generally, one could assume a non-compact group acting \emph{properly}, cf. \cite{P61}. \begin{Def}\label{shape:def} A complete connected finite-dimensional Riemannian manifold $M$ with geodesic distance $d_M$ on which a compact Lie group $G$ acts isometrically from the left is called a \emph{pre-shape space}. Moreover the canonical quotient $$\pi: M \to Q:= M/G = \{[p]: p \in M\}\mbox{ with the \emph{orbit} } [p] = \{gp: g\in G\}\,,$$ is called a \emph{shape space}. \end{Def} As a consequence of the isometric action we have that $d_M(gp,{p'}) = d(p,g^{-1}{p'})$ for all $p,{p'}\in M$, $g\in G$. For $p,{p'}\in M$ we say that $p$ is in \emph{optimal position} to ${p'}$ if $d_M(p,{p'}) =\min_{g\in G}d_M(gp,{p'})$, the minimum is attained since $G$ is compact. As is well known (e.g. \citep[p. 179]{Bre72}) there is an open and dense submanifold $M^*$ of $M$ such that the canonical quotient $Q^* = M^*/G$ restricted to $M^*$ carries a natural manifold structure also being open and dense in $Q$. Elements in $M^*$ and $Q^*$, respectively, are called \emph{regular}, the complementary elements are \emph{singular}; $Q^*$ is the \emph{manifold part} of $Q$. \paragraph{Intrinsic means on shape spaces.} The canonical quotient distance $$d_Q([p],[{p'}]) :=\min_{g\in G}d_M(gp,{p'}) = \min_{g,h\in G}d_M(gp,h{p'})$$ is called \emph{intrinsic distance} and the corresponding $d_Q$-Fr\'echet mean sets are called \emph{intrinsic means}. Note that the intrinsic distance on $Q^*$ is equal to the canonical geodesic distance. \paragraph{Ziezold and Procrustean means.} Now, assume that we have an embedding with orthogonal projection $\Phi: \mathbb R^D \to M\subset\mathbb R^D$ as above. If the action of $G$ is isometric w.r.t. the extrinsic metric, i.e. if $\|gp-g{p'}\| = \|p-{p'}\|$ for all $p,{p'}\in M$ and $g\in G$ then call \begin{eqnarray*} \rho^{(z)}_Q([p],[{p'}])&:=&\min_{g\in G}\|gp-{p'}\|\mbox{~~~and~~~}\\ \rho^{(p)}_Q([p],[{p'}])&:=&\min_{\footnotesize\begin{array}{l}g\in G,~ gp \mbox{ in}\\ \mbox{opt. pos. to }{p'}\end{array}}\|d\Phi_{p'}(gp-{p'})\| \end{eqnarray*} the \emph{Ziezold distance} and the \emph{Procrustean distance} on $Q$, respectively. Call the corresponding population and sample Fr\'echet $\rho^{(z)}_Q$-means, respectively, the sets of population and sample \emph{Ziezold means}, respectively. Similarly, call the corresponding population and sample Fr\'echet $\rho^{(p)}_Q$-means, respectively, the sets of population and sample \emph{Procrustean means}, respectively. We say that \emph{optimal positioning is invariant} if for all $p,{p'} \in M$ and $g^*\in G$, $$d_M(g^*p,{p'}) = \min_{g\in G}d_M(gp,{p'}) \Leftrightarrow \|g^*p-{p'}\| = \min_{g\in G}\|gp-{p'}\|\,.$$ \begin{Rm}\label{kendall_book:rm} Indeed for $Q=\Sigma_m^k, R\Sigma_m^k$, optimal positioning is invariant (cf. \citep[p. 206]{KBCL99}), Procrustean means coincide with means of \emph{general Procrustes analysis} introduced by \cite{Gow} and Ziezold means coincide with means as introduced by \cite{Z94} for $\Sigma_2^k$. Moreover for $\Sigma_2^k$, Procrustean means agree with extrinsic means w.r.t. the Veronese-Whitney embedding, cf. \cite{BP03} and Section \ref{ext-means:scn}. \end{Rm} Procrustean means on $\Sigma_m^k$ are also called \emph{full Procrustes means} in the literature to distinguish them from \emph{partial Procrustes means} on the \emph{size-and-shape spaces} not further discussed here (e.g. \cite{DM98}). We only note that partial Procrustes means are identical to the respective intrinsic, Procrustean and Ziezold means which on size-and-shape spaces, all agree with one another. \section{Horizontal Lifting and Manifold Stability}\label{convex:scn} In this section we derive a measurable horizontal lifting and the stability theorem underlying Theorem \ref{Kendall_mean_dim}. To this end we first recall how a shape space is made up from manifold strata of varying dimensions. Unless otherwise referenced, we use basic terminology that can be found in any standard textbook on differential geometry, e.g. \cite{KN63,KN69}. For the results derived here we assume that the shape space is a quotient modulo a compact group. We note that these results remain valid in the more general case of a non-compact group acting \emph{properly}, cf. \cite{P61}. \subsection{Preliminaries}\label{prelims:scn} Assume that $Q=M/G$ is a shape space as in Definition \ref{shape:def}. $T_pM$ is the tangent space of $M$ at $p\in M$ and $\exp_p$ denotes the \emph{Riemannian exponential} at $p$. Recall that on a Riemannian manifold the \emph{cut locus} $C(p)$ of $p$ comprises all points $q$ such that the extension of a length minimizing geodesic joining $p$ with $q$ is no longer minimizing beyond $q$. In consequence, on a complete and connected manifold $M$ we have for every $p'\in M$ that there is $v' \in T_pM$ such that $p'= \exp_p v'$ while $v' = \exp^{-1}_pp'$ of minimal modulus is uniquely determined as long as $p'\in M \setminus C(p)$. It is well known that the cut locus has measure zero in the sense that its image in any local chart has Lebesgue measure zero. From now on we call the cut locus the \emph{manifold cut locus} in order to distinguish it from the \emph{quotient cut locus} $C^{quot}(q)$ of $q\in Q$ which we define as $C^{quot}(q) := \{[p']: p'\in C(p) \mbox{ is in optimal position to some } p\in q\}\,.$ Due to the isometric action we have for any $p\in q$ that \begin{eqnarray}\label{quot-cut-loc:eq} C^{quot}(q) &=& \{[p']: p'\in C(p) \mbox{ is in optimal position to } p\}\subset \pi\big(C(p)\big). \end{eqnarray} The following Lemma teaches that in general, the projection of the manifold cut locus, the manifold cut locus of the manifold part $Q^*$ and the quotient cut locus are different. In particular, quotient cut loci are void in the special case of Kendall's shape spaces. \begin{Lem}\label{cut_locus:lem}$C(q)\neq \emptyset$ for every $q\in \Sigma_2^k$ while $C^{quot}(q)=\emptyset\mbox{ for all }q\in \Sigma_m^k\,.$ Similarly $C^{quot}(q)=\emptyset\mbox{ for all }q\in R\Sigma_m^k\,.$\end{Lem} \begin{proof} The first assertion follows from the fact that $\Sigma_2^k$ is a compact manifold. For the second assertion consider $[p]\in \Sigma_m^k$. Since $C(p) = \{-p\}$ for $p\in S_m^k$ and $[p]=[-p]$ for even $m$ as well as for odd $m$ if $p$ is not regular, and, since $p,-p$ are not in optimal position for odd $m$ if $p$ is regular, we have that $C^{quot}([p])=\emptyset$. The third assertion follows from the fact that $[p]=[-p]$ for all $[p] \in R\Sigma_m^k$. \end{proof} Next we collect consequences of the isometric Lie group action, see \cite{Bre72}. \begin{enumerate} \item[(A)] With the \emph{isotropy group} $I_p=\{g\in G: gp =p\}$ for $p\in M$, every orbit carries the natural structure of a coset space $[p] \cong G/I_p$. Moreover, $p' \in M$ is of \emph{orbit type} $(G/I_p)$ if $I_{p'} = gI_pg^{-1}=I_{gp}$ for a suitable $g\in G$. If $I_p \subset I_{gp'}$ for suitable $g\in G$ then $p'$ is of \emph{lower orbit type} than $p$ and $p$ is of \emph{higher orbit type} than $p'$. \item[(B)] The pre-shapes of equal orbit type $M^{(I_p)} := \{p'\in M:\mbox{ $p'$ is of orbit type }\linebreak (G/I_p)\}$ and the corresponding shapes $Q^{(I_p)}:=\{[p']:p'\in M^{(I_p)}\}$ are manifolds in $M$ and $Q$, respectively. Moreover, for $q\in Q$ denote by $Q^{(q)}$ the shapes of higher orbit type. \item[(C)]The orthogonal complement $H_pM$ in $T_pM$ of the tangent space $T_p[p]$ along the orbit is called the \emph{horizontal space}: $T_pM = T_p[p]\oplus H_pM$. \item[(D)] The \emph{Slice Theorem} states that every $p\in M$ has a \emph{tubular neighborhood} $[p]\subset U \subset M$ such that with a suitable subset $D\subset H_pM$ the \emph{twisted product} $\exp_p D\times_{I_p} G$ is diffeomorphic with $U$. Here, the twisted product is the natural topological quotient of the product space $\exp_p D\times G$ modulo the equivalence $$ (\exp_pv,g) \sim_{I_p} (\exp_pv', g')~\Leftrightarrow \exists h \in I_p\mbox{ such that } v'=dh v,~g'=gh^{-1}\,.$$ We then say that the tubular neighborhood $U$ \emph{admits a slice} $\exp_p D$ \emph{via} $U\cong \exp_p D\times_{I_p}G$ \item[(E)] Every $p\in M$ has a tubular neighborhood $U$ of $p$ that admits a slice $\exp_p D$ such that every $p' \in \exp_pD$ is in optimal position to $p$. Moreover, for any tubular neighborhood $U$ admitting a slice $\exp_p D$, all points $p'\in U$ are of orbit type higher than or equal to the orbit type of $p$ and only finitely many orbit types occur in $U$. If $p$ is regular, i.e. of maximal orbit type, then the product is trivial: $\exp_p D \times_{I_p} G \cong \exp_p D \times G/I_p$. \end{enumerate} Finally let us extend the following uniqueness property for the intrinsic distance to the Ziezold distance. The differential of the mapping $f^{p'}_{int} : M\setminus C(p') \to [0,\infty)$ defined by $f^{p'}_{int}(p) = d_M(p,\exp_pp')^2$ is given by $d f^{p'}_{int}(p) = -2v$ with $v=\exp_p^{-1}p'$ (cf. \citep[p. 110]{KN69}, \cite{Ka77}). Hence, we have for $p_1,p_2 \in M\setminus C(p)$ that \begin{eqnarray}\label{intr-inj:cond} d f^{p_1}_{int}(p)&=& df^{p_2}_{int}(p) ~\Leftrightarrow~ p_1=p_2\,. \end{eqnarray} In view of the extrinsic distance let $f^{p'}_{ext} : M\setminus C(p') \to [0,\infty)$ be defined by $f^{p'}_{ext}(p)=\|p-p'\|^2=\|p-\exp_p(\exp_p^{-1}p')\|^2$. Mimicking (\ref{intr-inj:cond}) introduce the following condition \begin{eqnarray}\label{Ziez-inj:cond} d f^{p_1}_{ext}(p) = d f^{p_2}_{ext}(p) &\Leftrightarrow& p_1=p_2 \end{eqnarray} for $p_1,p_2 \in M\setminus C(p)$. \begin{Rm}\label{half-sphere:rm} (\ref{Ziez-inj:cond}) is valid on closed half spheres since on the unit sphere $$ df_{ext}^{p'}(p) = -2\,\frac{v}{\|v\|}\,\sin(\|v\|)\mbox{ with } v=\exp_p^{-1}p'\,.$$ \end{Rm} \subsection{A Measurable Horizontal Lift} In order to establish the stability of means in Theorem \ref{population_mean_iso_grp:thm} in the following Section \ref{population_convex:scn}, here we lift a random shape $X$ from $Q$ horizontally to a random pre-shape $Y$ on $M$. In order to do so we need to guarantee the measurability of the horizontal lift in Theorem \ref{glob_hor_measurable_lift:thm} below, the proof of which can be found in the appendix. Before continuing, let us consider a simple example for illustration. Suppose that $G=S^1\subset \mathbb C$ acts on $M=\mathbb C$ by complex scalar multiplication. Then $[0,\infty) \cong Q=M/G$ having the two orbit types $(S^1/I_0) = \{1\}$ and $(S^1/I_1) = S^1$ gives rise to $Q^{(I_0)} = \{0\}$ and $Q^{(I_1)} = (0,\infty)$. Obviously, $M$ admits a global slice via the polar decomposition $[0,\infty) \times_{S^1} S^1 =\{0\} \cup \big((0,\infty)\times S^1\big)\cong M$ about $0\in M$ (the Riemannian exponential is the identity if $T_0\mathbb C$ is identified with $\mathbb C$). Here, $X$ can be identified with its horizontal lift $Y$ to the global slice $[0,\infty) \subset M$. If, say, $X$ is uniformly distributed on $[1,2]$ then $\mathbb P\{X\in Q^{(I_1)} \}>0$. In this case the stability theorem states the obvious fact that $0 \in Q^{(I_0)}$ cannot be a mean of $X$. \begin{Def}\label{hor-lift:def} Call a measurable subset $L \subset M$ a \emph{measurable horizontal lift} of a measurable subset $R$ of $M/G$ \emph{in optimal position to $p \in M$} if \begin{enumerate} \item the canonical projection $L \to R\subset M/G$ is surjective, \item every $p' \in L$ is in optimal position to $p\in L$, \item every orbit $[p']$ of $p'\in L$ meets $L$ once. \end{enumerate} \end{Def} \begin{Th}\label{glob_hor_measurable_lift:thm} Let $p\in [p]\in Q$ and $A\subset Q$ countable. Then there is a measurable horizontal lift $L$ of $Q^{([p])} \cup A$ in optimal position to $p$. \end{Th} \begin{Th}\label{pop_int_mean_bottom_int_mean_top:Th} Assume that $X$ is a random shape on $Q$ and that there are $p\in M$ and $A\subset Q$ countable such that $X$ is supported by $\big(Q^{([p])} \cup A\big)\setminus C^{quot}([p])$. With a measurable horizontal lift $L$ of $\big(Q^{([p])}\cup A\big)\setminus C^{quot}([p])$ in optimal position to $p$ define the random element $Y$ on $L\subset M$ by $\pi\circ Y = X$. \begin{enumerate}\item[(i)] If $[p]$ is an intrinsic mean of $X$ on $Q$, then $p$ is an intrinsic mean of $Y$ on $M$ and $$\mathbb E(\exp^{-1}_pY)=0\,.$$ \item[(ii)] If $[p]$ is a Ziezold mean of $X$ on $Q$ and optimal positioning is invariant, then $p$ is an extrinsic mean of $Y$ on $M$ and $$\mathbb E\big( d f_{ext}^{Y}(p)\big)=0\,.$$ \item[(iii)] If $[p]$ is a Procrustean mean of $X$ on $Q$ and optimal positioning is invariant, then $p$ is a residual mean of $Y$ on $M$. \end{enumerate} \end{Th} \begin{proof} Suppose that $[p]$ is an intrinsic mean of $X$. If $p$ would not be an intrinsic mean of $Y$, there would some $M\ni p'\neq p$ leading to the contradiction \begin{eqnarray*} \mathbb E \big(d_Q([p'], X)^2\big) &=& \mathbb E \big(d_M(p',Y)^2\big) ~<~ \mathbb E \big(d_M(p,Y)^2\big) ~=~ \mathbb E \big(d_Q([p], X)^2\big)\,. \end{eqnarray*} Hence, $p$ is an intrinsic mean of $Y$. Replacing $d_Q$ by $\rho_Q^{(z)}$ and $d_M$ by the Euclidean distance gives the assertion for Ziezold and extrinsic means, respectively; and, using the Procrustean distance $\rho_Q^{(p)}$ on $Q$ as well as the residual distance $\rho^{(r)}_M$ on $M$ gives the assertion for Procrustean and residual means, respectively. For intrinsic means $p\in M$, the necessary condition $\mathbb E \big(\exp_{p}^{-1}Y\big) =0\,$ is developed in \citep[p. 110]{KN69}, cf. also \citep{Ka77} and \citep[p. 395]{KWS90} which yields the asserted equality in (i). By definition, the analog condition for an extrinsic mean $p\in M$ is $\mathbb E\big(d f^{Y}_{ext}(p)\big) =0$ which is the asserted equality in (ii) completing the proof. \end{proof} \begin{Rm}\label{Ziez-half-sphere:rm} Since the maximal intrinsic distance on $\Sigma_m^k$ and $R\Sigma_m^k$ is $$\frac{\pi}{2} = \max_{x,y\in S_m^k}\mathop{\min}_{g\in SO(m)}\arccos\big(\operatorname {tr}(gxy^T)\big)=\max_{x,y\in S_m^k}\mathop{\min}_{g\in O(m)}\arccos\big(\operatorname {tr}(gxy^T)\big)\,,$$ taking into account Remark \ref{half-sphere:rm}, condition (\ref{Ziez-inj:cond}) is satisfied for any horizontal lift in optimal position. \end{Rm} \subsection{Manifold Stability \label{population_convex:scn} The proof of the following central theorem is deferred to the appendix. \begin{Th}\label{population_mean_iso_grp:thm} Assume that $X$ is a random shape on $Q$, $p\in M$ and that $A\subset Q$ is countable such that $X$ is supported by $\big(Q^{([p])}\cup A\big)\setminus C^{quot}([p])$ and let $p'\in [p'] \in Q^{([p])}$. If ${\operatorname {\mathbb P}} \{X\in Q^{(I_{p'})}\} \neq 0$ and if either $[p]$ is \begin{enumerate} \item[(i)] an intrinsic mean of $X$ or \item[(ii)] a Ziezold mean of $X$ while optimal positioning is invariant and (\ref{Ziez-inj:cond}) is valid, \end{enumerate} then $p'$ is of lower orbit type than $p$. \end{Th} We have at once the following Corollary. \begin{Cor}[Manifold Stability Theorem]\label{pop_mean_reg:cor} Suppose that $X$ is a random shape on $Q$ that is supported by $Q\setminus C^{quot}([p])$ for some $[p] \in Q$ assuming the manifold part $Q^*$ with non-zero probability and having at most countably many point masses on the singular part. Then $[p]$ is regular if it is an intrinsic mean of $X$, or if it is a Ziezold mean, optimal positioning is invariant and (\ref{Ziez-inj:cond}) is valid. \end{Cor} Since $Q\setminus Q^{(q)}$ is a null set in $Q$ for every $q\in Q$ (cf. \citep[p. 184]{Bre72}) and so is $C^{quot}(q)$ -- by (\ref{quot-cut-loc:eq}) it is contained in the projection of a null set -- we have the following practical application. \begin{Cor}\label{cont_distr:cor} Suppose that a random shape on $Q$ is absolutely continuously distributed w.r.t. the projection of the Riemannian volume on $M$. Then intrinsic and Ziezold population means are regular; the latter if optimal positioning is invariant and (\ref{Ziez-inj:cond}) is valid. And, intrinsic and Ziezold sample means are a.s. regular. \end{Cor} \subsection[Non-Stability for Procrustean Means]{An Example for Non-Stability of Procrustean Means}\label{Procrustes-convex:scn} Consider a random configuration $Z\in F_3^4$ assuming the collinear quadrangle $q_1$ with probability $2/3$ and the planar quadrangle $q_2$ with probability $1/3$ where $$ q_1=\left(\begin{array}{cccc}1&-1&0&0\\0&0&0&0\\0&0&0&0\end{array}\right),~~ q_2= \left(\begin{array}{cccc}1&1&-2&0\\\frac{1}{\sqrt{2}}&\frac{1}{\sqrt{2}}&\frac{1}{\sqrt{2}}&-\,\frac{3}{\sqrt{2}}\\0&0&0&0\end{array}\right)\,.$$ Corresponding pre-shapes in optimal position w.r.t. the action of $SO(3)$ and $O(3)$ are given by $$ p_1=\left(\begin{array}{ccc}1&0&0\\0&0&0\\0&0&0\end{array}\right),~~ p_2=\frac{1}{\sqrt{2}}\, \left(\begin{array}{ccc}0&1&0\\0&0&1\\0&0&0\end{array}\right)\,.$$ Note that $[p_2]$ has regular shape in $(\Sigma_3^4)^*$. The full Procrustes mean of $[Z]\in \Sigma_3^4$ is easily computed to have the singular shape $[p_1]\in \Sigma_3^4\setminus (\Sigma_3^4)^*$, see Figure \ref{different_means_fig} as well as Examples \ref{different_means_ex} and Section \ref{blindness:scn}. Cf. also Remark \ref{sharp:rm}. \section{Extrinsic Means for Kendall's (Reflection) Shape Spaces}\label{ext-means:scn} Let us recall the well known \emph{Veronese-Whitney} embedding for Kendall's planar shape spaces $\Sigma_2^k$. Identify $F_2^k$ with $\mathbb C^{k-1}\setminus \{0\}$ such that every landmark column corresponds to a complex number. This means in particular that $z\in \mathbb C^{k-1}$ is a complex row(!)-vector. With the Hermitian conjugate $a^* = (\overline{a_{kj}})$ of a complex matrix $a=(a_{jk})$ the pre-shape sphere $S_2^k$ is identified with $\{z\in \mathbb C^{k-1}: zz^*=1\}$ on which $SO(2)$ identified with $S^1=\{\lambda \in\mathbb C: |\lambda|=1\}$ acts by complex scalar multiplication. Then the well known Hopf-Fibration mapping to complex projective space gives $\Sigma_2^k=S_2^k/S^1=\mathbb CP^{k-2}$. Moreover, denoting with $M(k-1,k-1,\mathbb C)$ all complex $(k-1)\times (k-1)$ matrices, the Veronese-Whitney embedding is given by \begin{eqnarray*} S_2^k/S^1 &\to& \{a \in M(k-1,k-1,\mathbb C): a^*=a\}\,,~~ ~[z] ~\mapsto~ z^*z\,. \end{eqnarray*} \begin{Rm}\label{VW-emb-iso:rm} The Procrustean metric of $\Sigma_2^k$ is isometric with the Euclidean metric of $M(k-1,k-1,\mathbb C)$ since we have $\langle z,w \rangle = \operatorname {Re}(zw^*)$ for $z,w\in S_2^k$ and hence, $d^{(p)}_{\Sigma_2^k}([z],[w]) = \sqrt{1-wz^*zw^*} = \|w^*w-z^*z\|/\sqrt{2}$. \end{Rm} The idea of the Veronese-Whitney embedding can be carried to the general case of shapes of arbitrary dimension $m\geq 2$. Even though the embedding given below is apt only for reflection shape space it can be applied to practical situations in similarity shape analysis whenever the geometrical objects considered have a common orientation. As above, the number $k$ of landmarks is essential and will be considered fixed throughout this section; the dimension $1\leq m < k$, however, is lost in the embedding and needs to be retrieved by projection. To this end recall the embedding of $S_j^k$ in $S_m^k$ $(1\leq j \leq m)$ in (\ref{pre-shape-emb:eq}) which gives rise to a canonical embedding of $R\Sigma_j^m$ in $R\Sigma_m^k$. Moreover, consider the strata $$(R\Sigma_m^k)^j := \{[x]\in R\Sigma_m^k: \operatorname {rank}(x)=j\},~~(\Sigma_m^k)^{j}:=\{[x] \in \Sigma_m^k: \operatorname {rank}(x)=j\}$$ for $j=1,\ldots,m$, each of which carries a canonical manifold structure; due to the above embedding, $(R\Sigma_m^k)^j$ will be identified with $(R\Sigma_j^k)^j$ such that $$ R\Sigma_m^k ~=~ \bigcup_{j=1}^m (R\Sigma_j^k)^j\,,$$ and $(R\Sigma_m^k)^j$ with $(\Sigma_m^k)^j$ in case of $j<m$. At this point we note that $SO(m)$ is connected, while $O(m)$ is not; and the consequences for the respective manifold parts, i.e. points of maximal orbit type: \begin{eqnarray}\label{mf-part:eq} (\Sigma_m^k)^* &=& (\Sigma_m^k)^{m-1}\cup (\Sigma_m^k)^m\,,\qua (R\Sigma_m^k)^* ~=~ (R\Sigma_m^k)^{m}\,. \end{eqnarray} Similarly, we have a stratifiction $${\cal P} := \left\{a\in M(k-1,k-1): a=a^T\geq 0, \operatorname {tr}(a)=1\right\} ~=~ \bigcup_{j=1}^{k-1}{\cal P}^j$$ of a compact flat convex space ${\cal P}$ with non-flat manifolds ${\cal P}^j :=\{a\in {\cal P}: \operatorname {rank}(a) = j\}~ (j=1,\ldots,k-1)\,,$ all embedded in $M(k-1,k-1)$. The \emph{Schoenberg map} $\mathfrak{s} : R\Sigma_m^k \to {\cal P}$ is then defined on each stratum by $$ \begin{array}{rcl}\mathfrak{s}|_{(R\Sigma_m^k)^j}=:\mathfrak{s}^j: (R\Sigma_m^k)^j &\to& {\cal P}^j \,,~~[x]~\mapsto~x^Tx\end{array}\,.$$ For $x\in S_j^k$ recall the tangent space decomposition $T_xS_j^k = T_{x}[x]\oplus H_xS_j^k$ into the \emph{vertical tangent space} along the orbit $[x]$ and its orthogonal complement the \emph{horizontal tangent space}. For $x\in (S_j^k)^j$ identify canonically (cf. \citep[p. 109]{KBCL99}): $$ T_{[x]} (R\Sigma_j^k)^j\cong H_xS_j^k = \{w \in M(j,k-1): \operatorname {tr}(wx^T) = 0, wx^T=xw^T\}\,.$$ Then the assertion of the following Theorem condenses results of \cite{BandPat05}, cf. also \cite{DKLW08}. \begin{Th} Each $\mathfrak{s}^j$ is a diffeomorphism with inverse $(\mathfrak{s}^j)^{-1}(a) = [(\sqrt{\lambda} u^T)_1^j]$ where $a = u\lambda u^T$ with $u \in O(k-1)$, $\lambda=\operatorname {diag}(\lambda_1,\ldots,\lambda_m)$, and $0=\lambda_{j+1} =\ldots=\lambda_{k-1}$ in case of $j<k-1$. Here, $(a)_1^j$ denotes the matrix obtained from taking only the first $j$ rows from $a$. For $x\in S_j^k$ and $w\in H_xS_j^k\cong T_{[x]}(R\Sigma_j^k)^j$ the derivative is given by $$d(\mathfrak{s}^j)_{[x]}w = x^Tw+w^Tx\,.$$ \end{Th} \begin{Rm}\label{Schoenberg-emb-non-iso:rm} In contrast to the Veronese-Whitney embedding, the Schoenberg embedding is not isometric as the example of $$ x = \left(\begin{array}{cc}\cos \phi&0\\0&\sin\phi\end{array}\right)\,,~~ w_1 = \left(\begin{array}{cc}\sin \phi&0\\0&-\cos\phi\end{array}\right),~~ w_2 = \left(\begin{array}{cc}0&\cos \phi\\\sin\phi&0\end{array}\right),~~$$ teaches: $\|x^Tw_1+w_1^Tx\| = \sqrt{2}\, 2|\cos\phi\sin\phi|,~~ \|x^Tw_2+w_2^Tx\| =\sqrt{2}\,.$ \end{Rm} Since ${\cal P}$ is bounded, convex and Euclidean, the classical expectation $\mathbb E(X^TX)\in {\cal P}^j$ for some $1\leq j\leq k-1$ of the Schoenberg image $X^TX$ of an arbitrary random reflection shape $[X]\in R\Sigma_m^k$ is well defined. Then we have at once the following relation between the rank of the Euclidean mean and increasing sample size. \begin{Th}\label{dim-Schoenberg-mean:thm} Suppose that a random reflection shape $[X]\in R\Sigma_m^k$ is distributed absolutely continuous w.r.t. the projection of the spherical volume on $S_m^k$. Then $$\mathbb E(X^TX)\in {\cal P}^{k-1}\mbox{ and~~~} \frac{1}{n}\sum_{i=1}^n X_i^TX_i\in {\cal P}^{\min( nm , k-1)}~ a.s.$$ for every i.i.d. sample $X_1,\ldots,X_n\sim X$. \end{Th} Hence, in stastical settings involving a higher number of landmarks, a sufficiently well behaved projection of a high rank Euclidean mean onto lower rank ${\cal P}^r\cong (\Sigma_r^k)^r$, usually $m=r$, is to be employed, giving at once a mean shape satisfying strong consistency and a CLT. Here, unlike to intrinsic or Procrustes analysis, the dimension $r$ chosen is crucial for the dimensionality of the mean obtained. The orthogonal projection $$\begin{array}{rcl}\phi^r : \bigcup_{i=r}^{k-1} {\cal P}^i &\to& {\cal P}^r \,,~~ a~\mapsto~\operatorname {argmin}_{b \in {\cal P}^r} \operatorname {tr}\big((a - b)^2\big)\end{array}$$ giving the set of \emph{extrinsic Schoenberg means} has been computed by \cite{B08}: \begin{Th}\label{og-proj-sb:thm} For $1\leq r\leq k-1$, $a = u\lambda u^T \in {\cal P}$ with $u \in O(k-1)$, $\lambda=\operatorname {diag}(\lambda_1,\ldots,\lambda_m)$, $\lambda_{1}\geq \ldots\geq \lambda_{k-1}$ and $\lambda_r>0$ the orthogonal projection onto ${\cal P}^r$ is given by $ \phi^r(a) = u \mu u^T$ with $\mu = \operatorname {diag}(\mu_1,\ldots,\mu_r,0,\ldots,0)$, $$\mu_i = \lambda_i +\frac{1}{r} - \overline{\lambda}_r\,(i=1,\ldots,r)$$ and $\overline{\lambda}_r = \frac{1}{r}\sum_{i=1}^r\lambda_i \leq \frac{1}{r}$ which is uniquely determined if and only if $\lambda_{r}>\lambda_{r+1}$. \end{Th} With the notation of Theorem \ref{og-proj-sb:thm}, a non-orthogonal \emph{central projection} $\psi^r(a) = u \nu u^T$ equally well and uniquely determined has been proposed by \cite{DKLW08} with $$\nu = \operatorname {diag}(\nu_1,\ldots,\nu_r,0,\ldots,0),~~ \nu_i = \frac{\lambda_i}{r\overline{\lambda}_r}~~(i=1,\ldots,r)\,.$$ Orthogonal and central projection are depicted in Figure \ref{Schoenberg_proj_fig}. \begin{figure} \begin{minipage}{0.51\textwidth} \includegraphics[angle=0,width=1\textwidth]{project_lambda_plane1.eps} \end{minipage} \begin{minipage}{0.48\textwidth} \caption{\it Projections (if existent) of two points (crosses) in the $\lambda$-plane to the open line segment $\Lambda=\{(\lambda_1,\lambda_2): \lambda_1+\lambda_2=1, \lambda_1,\lambda_2>0\}$. The dotted line gives the central projections (denoted by stars) which is well defined for all symmetric, positive definite matrices (corresponding to the first open quadrant), the dashed line gives the orthogonal projection (circle) which is well defined in the triangle below $\Lambda$ (corresponding to ${\cal P}$) and above $\Lambda$ in an open strip. In particular it exists not for the right point.\label{Schoenberg_proj_fig}} \end{minipage} \end{figure} \section{Local Effects of Curvature}\label{local:scn} In this section we assume that a manifold stratum $M$ supporting a random element $X$ is isometrically embedded in a Euclidean space $\mathbb R^D$ of dimension $D>0$. With the orthogonal projection $\Phi :\mathbb R^D \to M$ from Section \ref{Frechet-rho-means:scn} and the Riemannian exponential $\exp_p$ of $M$ at $p$ we have the $$\begin{array}{lcl} \mbox{\emph{intrinsic tangent space coordinate }}&& \exp_p^{-1}X\mbox{ and the}\\ \mbox{\emph{residual tangent space coordinate }}&& d\Phi_p(X-p)\,, \end{array}$$ respectively, of $X$ at $p$, if existent. \subsection{Finite Power of Tests and Tangent Space Coordinates}\label{tangent:scn} With the above setup, assume that $\mu\in M$ is a unique mean of $X$. Moreover, we assume that $M$ is curved near $\mu$, i.e. that there is $c\in \mathbb R^D$, the center of the osculatory circle touching the geodesic segment in $M$ from $X$ to $\mu$ at $\mu$ with radius $r$. If $X_r$ is the orthogonal projection of $X$ to that circle, then $X=X_r + O(\|X-\mu\|^3)$. Moreover, with \begin{eqnarray*} \cos \alpha &=& \left\langle \frac{X-c}{\|X-c\|}, \frac{\mu-c}{r}\right\rangle ~=~ \frac{1}{r^2}\,\langle X_r-c,\mu-c\rangle + O(\|X-\mu\|^3) \end{eqnarray*} we have the residual tangent space coordinate $$ v = X-c- \frac{\mu-c}{r}\,\|X-c\|\,\cos \alpha = X_r-c -(\mu-c)\cos \alpha+ O(\|X-\mu\|^3)$$ having squared length $\|v\|^2 =r^2\sin\alpha^2 + O(\|X-\mu\|^3)$. By isometry of the embedding, the intrinsic tangent space coordinate is given by $$\exp^{-1}_{\mu} X = \frac{r\alpha}{\|v\|}\, v + O(\|X-\mu\|^3)\,.$$ With the component $$ \nu = \mu-c - \|X-c\| \frac{\mu-c}{r}\,\cos \alpha = (\mu -c) (1-\cos \alpha) + O(\|X-\mu\|^3) $$ of $X$ normal to the above mentioned geodesic segment of squared length $\|\nu\|^2 = r^2(1-\cos\alpha)^2+O(\|X-\mu\|^3)$, we obtain $ \|\exp^{-1}_{\mu} X\|^2 ~=~ \|v\|^2 + \|\nu\|^2+ O(\|X-\mu\|^3)\,, $ since \begin{eqnarray*} (1-\cos\alpha)^2 + \sin^2\alpha &=& 2(1-\cos\alpha) ~=~\alpha^2 + 2\frac{\alpha^4}{4!} + \cdots \end{eqnarray*} and $\alpha = O(\|X-\mu\|)$. In consequence we have \begin{Rm}\label{use_is_ext_not_intr_coord_rm} In approximation, the variation of intrinsic tangent space coordinates is the sum of the variation $\|v\|^2$ of residual tangent space coordinates and the variation normal to it. In particular, for spheres $$ \|\exp^{-1}_{\mu} X\|^2 ~\geq~ \|v\|^2 + \|\nu\|^2\,.$$ Since the variation in normal space is irrelevant for a two-sample test for equality of means, say, a higher power for tests based on intrinsic means can be expected when solely residual tangent space coordinates obtained from an isometric embedding are used rather than intrinsic tangent space coordinates. Note that the natural tangent space coordinates for Ziezold means are residual. \end{Rm} A simulated classification example in Section \ref{class_data_simulations_scn} illustrates this effect. \subsection{The $1:3$ - Property for Spherical and Kendall Shape Means}\label{1:3:scn} In this section $M=S^{D-1}\subset \mathbb R^D$ is the $(D-1)$-dimensional unit-hypersphere embedded isometrically in Euclidean $D$-dimensional space. The orthogonal projection $\Phi: \mathbb R^D \to S^{D-1}: p \to \frac{p}{\|p\|}$ is well defined except for the origin $p=0$, and the normal space at $p\in S^{D-1}$ is spanned by $p$ itself. In consequence, a random point $X$ on $S^{D-1}$ has $$d\Phi_p(X-p) = X -p \cos \alpha,~~\exp^{-1}_p(X) = \left\{\begin{array}{ll}\frac{\alpha}{\sin \alpha}~ d\Phi_p(X-p)&\mbox{ for }X\neq p\\0&\mbox{ for }X=p\end{array}\right.$$ as residual and intrinsic, resp., tangent space coordinate at $-X\neq p\in S^{D-1}$ where $\cos\alpha = \langle X,p\rangle$, $\alpha \in [0,\pi)$. \begin{Th}\label{unique-intr-mean:th} If $X$ a.s. is contained in an open half sphere, it has a unique intrinsic mean which is assumed in the interior of that half sphere. \end{Th} \begin{proof} Below, we show that every intrinsic mean necessarily lies within the interior of the half sphere. Then, \citep[Theorem 7.3]{KWS90} yields uniqueness. W.l.o.g. let $X = (\sin\phi, x_2,\ldots, x_n)$ such that ${\operatorname {\mathbb P}}\{\sin\phi \leq 0\} = 0 =1- {\operatorname {\mathbb P}}\{\sin\phi >0\}$ and assume that $p = (\sin \psi, p_2,\ldots, p_n)\in S^{D-1}$ is an intrinsic mean, $-\pi/2\leq \phi,\psi \leq \pi/2$. Moreover let $p' = (\sin (|\psi|), p_2,\ldots, p_n)$. Since \begin{eqnarray*} \operatorname {\mathbb E}\left(\|\exp^{-1}_p(X)\|^2\right) &=& \operatorname {\mathbb E}\left(\arccos^2 \langle p,X\rangle\right)\\ &=&\operatorname {\mathbb E}\left(\arccos^2 \left(\sin\psi\sin\phi + \sum_{j=2}^np_jx_j\right)\right)\\ &\geq& \operatorname {\mathbb E}\left(\|\exp^{-1}_{p'}(X)\|^2\right) \end{eqnarray*} with equality if and only if $\sin |\psi| =\sin \psi$, this can only happen for $\sin\psi\geq 0$. Now, suppose that $p=(0,p_2,\ldots,p_n)$ is an intrinsic mean. For small deterministic $\psi \geq 0 $ consider $p(\psi) = (\sin \psi,p_1\cos\psi,\ldots,p_n\cos\psi)$. Then \begin{eqnarray*} \operatorname {\mathbb E}\left(\|\exp^{-1}_{p(\psi)}(X)\|^2\right) &=&\operatorname {\mathbb E}\left(\arccos^2 \left(\sin\psi\sin\phi + \cos \psi \sum_{j=2}^np_jx_j\right)\right)\\ &=& \operatorname {\mathbb E}\left(\|\exp^{-1}_{p}(X)\|^2\right) - C_1\psi + O(\psi^2) \end{eqnarray*} with $C_1 >0$ since ${\operatorname {\mathbb P}}\{\sin\phi >0\}>0$. In consequence, $p$ cannot be an intrinsic mean. Hence, we have shown that every intrinsic mean is contained in the interior of the half sphere. \end{proof} \begin{Rm}\label{Karcher-means:rm} For the special case of spheres, this is a simple proof for the general theorem recently established by \cite{Afsari10} which extends results of \cite{Ka77,KWS90} and \cite{L01,L04}, stating that the intrinsic mean on a general manifold is unique if among others the support of the distribution is contained in a geodesic half ball. \end{Rm} The following theorem characterizes the three spherical means. \begin{Th}\label{conditions_for_spherical_means_th} Let $X$ be a random point on $S^{D-1}$. Then $x^{(e)} \in S^{D-1}$ is the unique extrinsic mean if and only if the Euclidean mean $\mathbb E(X) = \int_{S^{D-1}} X\,d\operatorname {\mathbb P}_X$ is non-zero. In that case $$\lambda^{(e)} x^{(e)} = \mathbb E(X)\,$$ with $\lambda^{(e)} = \|\mathbb E(X)\|>0$. Moreover, there are suitable $\lambda^{(r)}>0$ and $\lambda^{(i)}>0$ such that every residual mean $x^{(r)} \in S^{D-1}$ satisfies $$\lambda^{(r)} x^{(r)} = \mathbb E\big(\langle X,x^{(r)}\rangle\, X\big)\,,$$ and every intrinsic mean $x^{(i)} \in S^{D-1}$ satisfies $$\lambda^{(i)} x^{(i)} = \mathbb E\left(\frac{\arccos\langle X,x^{(i)}\rangle}{\sqrt{1-\langle X,x^{(i)}\rangle^2 }}\, X\right)\,.$$ In the last case we additionally require that $\operatorname {\mathbb E}\left(\frac{\arccos\langle X,x^{(i)}\rangle}{\sqrt{1-\langle X,x^{(i)}\rangle^2 }}\, \langle X,x^{(i)}\rangle\right)> 0$ which is in particular the case if $X$ is a.s. contained in an open half sphere. \end{Th} \begin{proof} The assertions for the extrinsic mean are well known from \cite{HLR96}. The second assertion for residual means follows from minimization of $$ \int_{S^{D-1}} \|p-\langle p,x\rangle x\|^2\,d\operatorname {\mathbb P}_X(p) = 1 - \int_{S^{D-1}} \langle p,x\rangle^2\,d\operatorname {\mathbb P}_X(p)$$ with respect to $x\in \mathbb R^D$ under the constraining condition $\|x\|=1$. Using a Lagrange ansatz this leads to the necessary condition $$ \int_{S^{D-1}} \langle p,x\rangle\, p\,d\operatorname {\mathbb P}_X(p) = \lambda x$$ with a Lagrange multiplier $\lambda$ of value $\operatorname {\mathbb E}(\langle X,x\rangle^2)$ which is positive unless $X$ is supported by the hypersphere orthogonal to $x$. In that case, by Proposition \ref{res_means_sphere:prop}, $x$ cannot be a residual mean of $X$, as every residual mean is as well contained in that hypersphere. Hence, we have $\lambda^{(r)}:=\lambda >0$. The Lagrange method applied to $$ \int_{S^{D-1}} \|\exp^{-1}_x(p)\|^2 \,d\operatorname {\mathbb P}_X(p) = \int_{S^{D-1}} \arccos^2(\langle p,x\rangle)\,d\operatorname {\mathbb P}_X(p) $$ taking into account Theorem \ref{unique-intr-mean:th}, insuring that $x^{(i)}$ is in the open half sphere that contatins $X$ a.s., yields the third assertion on the intrinsic mean. \end{proof} Recall that residual means are eigenvectors to the largest eigenvalue of the matrix $\mathbb E(XX^T)$. As such, they rather reflect the mode than the classical mean of a distribution: \begin{Ex}\label{different_means_ex} Consider $\gamma \in (0,\pi)$ and a random variable $X$ on the unit circle $\{e^{i\theta}: \theta \in [0,2\pi)\}$ which takes the value $1$ with probability $2/3$ and $e^{i\gamma}$ with probability $1/3$. Then, explicit computation gives the unique intrinsic and extrinsic mean as well as the two residual means $$ x^{(i)} = e^{i\frac{\gamma}{3}},\quad x^{(e)} = e^{i\arctan\frac{\sin \gamma}{2+\cos\gamma}},\quad x^{(r)} = \pm\, e^{i\frac{1}{2}\,\arctan\frac{\sin (2\gamma)}{2+\cos(2\gamma)}}\,.$$ Figure \ref{different_means_fig} shows the case $\gamma = \frac{\pi}{2}$. \end{Ex} \begin{figure} \begin{minipage}{0.5\textwidth} \includegraphics[width=0.95\textwidth]{different_means_S1.eps} \end{minipage} \begin{minipage}{0.49\textwidth} \caption{\it Means on a circle of a distribution taking the upper dotted value with probability $1/3$ and the lower right dotted value with probability $2/3$. The latter happens to be one of the two residual means.\label{different_means_fig}} \end{minipage} \end{figure} In contrast to Figure \ref{different_means_fig}, one may assume in many practical applications that the mutual distances of the unique intrinsic mean $x^{(i)}$, the unique extrinsic mean $x^{(e)}$ and the unique residual mean $x^{(r_0)}$ closer to $x^{(e)}$ are rather small, namely of the same order as the squared proximity of the modulus $\|\mathbb E(X)\|$ of the Euclidean mean to $1$, cf. Table \ref{1:3:tab}. We will use the following condition \begin{eqnarray}\label{reg_extr_mean_cond}\left.\begin{array}{rcl} \|x^{(e)} -x^{(r_0)}\|\,,~~ \|x^{(e)} -x^{(i)}\|&=& O\big((1-\|\operatorname {\mathbb E}(X)\|)^2\big) \end{array}\right. \end{eqnarray} with the \emph{concentration parameter} $1-\|\operatorname {\mathbb E}(X)\|$. \begin{Cor}\label{3_means_on_circle_cor} Under condition (\ref{reg_extr_mean_cond}), if all three means are unique, then the great circular segment between the residual mean $x^{(r_0)}$ closer to the extrinsic mean $x^{(e)}$ and the intrinsic mean $x^{(i)}$ is divided by the extrinsic mean in approximation by the ratio $1:3$: \begin{eqnarray* x^{(r_0)} &=& \frac{\|\operatorname {\mathbb E}(X)\|}{\lambda^{(r)}}\left(x^{(e)} - \frac{\mathbb E\big( \langle X-x^{(e)} ,X\rangle\,X\big)}{\|\operatorname {\mathbb E}(X)\|} + O\big((1-\|\operatorname {\mathbb E}(X)\|)^2\big)\right)\\ x^{(i)} &=& \frac{\|\operatorname {\mathbb E}(X)\|}{\lambda^{(i)}}\left(x^{(e)} + \frac{1}{3}\, \frac{\mathbb E\big( \langle X-x^{(e)} ,X\rangle\,X\big)}{\|\operatorname {\mathbb E}(X)\|} + O\big((1-\|\operatorname {\mathbb E}(X)\|)^2\big)\right)\, \end{eqnarray*} with $\lambda^{(i)}$ and $\lambda^{(r)}$ from Theorem \ref{conditions_for_spherical_means_th}. \end{Cor} \begin{proof} For any $x,p \in S^{D-1}$ decompose $ p - x = p - \langle x,p\rangle\, x - z(x,p) x $ with $z(x,p) = 1 - \langle x,p\rangle$, the length of the part of $p$ normal to the tangent space at $x$. Note that $\operatorname {\mathbb E}\big(z(x^{(e)},X\big) = 1-\|\operatorname {\mathbb E}(X)\|$. Now, under condition (\ref{reg_extr_mean_cond}), verify the first assertion using Theorem \ref{conditions_for_spherical_means_th}: \begin{eqnarray*}\label{approx_extr_res_mean}\nonumber x^{(r0)} &=&\frac{1}{\lambda^{(r)}}\left(\operatorname {\mathbb E}(X) - \operatorname {\mathbb E}\big(z(x^{(r_0)},X)X\big)\right)\,. \end{eqnarray*} On the other hand since \begin{eqnarray*} \frac{\arccos (1-z)}{\sqrt{1-(1-z)^2 }} &=& 1 + \frac{1}{3}\,z +\frac{2}{15}\,z^2 +\ldots \end{eqnarray*} we obtain with the same argument the second assertion \begin{eqnarray*} x^{(i)} &=& \frac{1}{\lambda^{(i)}} \operatorname {\mathbb E} \left(\frac{\arccos\langle X,x^{(i)}\rangle}{\sqrt{1-\langle X,x^{(i)}\rangle^2 }}\, X\right)\\ &=& \frac{1}{\lambda^{(i)}}\left(\operatorname {\mathbb E}(X)+ \frac{1}{3}\, \operatorname {\mathbb E}\left( z(x^{(i)},X)\,X\right) + \frac{2}{15}\, \operatorname {\mathbb E}\left( z(x^{(i)},X)^2\,X\right) +\ldots\right)\\ &=& \frac{\|\operatorname {\mathbb E}(X)\|}{\lambda^{(i)}}\left(x^{(e)} +\frac{1}{3\|\operatorname {\mathbb E}(X)\|}\, \operatorname {\mathbb E} \left(z(x^{(e)},X)\,X\right) + O\big((1-\|\operatorname {\mathbb E}(X)\|)^2\big)\right)\,. \end{eqnarray*} \end{proof} \begin{Rm} The tangent vector defining the great circle approximately connecting the three means is obtained from correcting with the expected normal component of any of the means. As numerical experiments show, this great circle is different from the first \emph{principal component geodesic} as defined in \cite{HZ06}. \end{Rm} Recall the following connection between top and quotient space means, cf. Theorem \ref{pop_int_mean_bottom_int_mean_top:Th}. \begin{Rm}\label{kend-mf-quot-mean:rm} Let $p \in S_m^k$ such that a random shape $X$ on $\Sigma_m^k$ is supported by $(\Sigma_m^k)^{([p])}\cup A$ with $A\subset \Sigma_m^k$ at most countable. By Lemma \ref{cut_locus:lem} and Theorem \ref{pop_int_mean_bottom_int_mean_top:Th}, $(\Sigma_m^k)^{([p])}\cup A$ admits a horizontal measurable lift $L\subset S_m^k$ in optimal position to $p\in S_m^k$. Define the random variable $Y$ on $L\subset S_m^k$ by $\pi \circ Y = X$. Then we have that \begin{enumerate} \item[] if $[p]$ is an intrinsic mean of $X$ then $p$ is an intrinsic mean of $Y$, \item[] if $[p]$ is a full Procrustean mean of $X$ then $p$ is a residual mean of $Y$, \item[] if $[p]$ is a Ziezold mean of $X$ then $p$ is an extrinsic mean of $Y$. \end{enumerate} \end{Rm} In consequence, Corollary \ref{3_means_on_circle_cor} extends at once to Kendall's shape spaces. \emph{Generalized geodesics} referred to below are an extension of the concept of geodesics to non-manifold shape spaces, cf. \cite{HHM07}. \begin{Cor}\label{3_means_on_kendall_cor} Suppose that a random shape $X$ on $\Sigma_m^k$ with unique intrinsic mean $\mu^{(i)}$, unique Ziezold mean $\mu^{(z)}$ and unique Procrustean mean $\mu^{(p)}$ is supported by $R=(\Sigma_m^k)^{(\mu^{(i)})}\cap(\Sigma_m^k)^{(\mu^{(z)})}\cap (\Sigma_m^k)^{(\mu^{(p)})}$. If the means are sufficiently close to each other in the sense of $$d_{\Sigma_m^k}(\mu^{(z)} -\mu^{(p)})\,,~~ d_{\Sigma_m^k}(\mu^{(z)} -\mu^{(i)})~=~ O\big((1-\|\operatorname {\mathbb E}(Y)\|)^2\big) $$ with the random pre-shape $Y$ on a horizontal lift $L$ of $R$ defined by $X = \pi \circ Y$, then the generalized geodesic segment between $\mu^{(i)}$ and $\mu^{(p)}$ is approximately divided by $\mu^{(z)}$ by the ratio $1:3$ with an error of order $O\big((1-\|\operatorname {\mathbb E}(Y)\|)^2\big)$. \end{Cor} \section{Examples: Exemplary Datasets and Simulations}\label{class_data_simulations_scn} All of the results of this section are based on datasets and simulations, i.e., all means considered are sample means. An R-package for the computation of all means including the poplar leaves data can be found under \cite{Hshapes}. \subsection{The $1:3$ property} In the first example we illustrate Corollary \ref{3_means_on_kendall_cor} on the basis of four classical data sets: \begin{description} \item[poplar leaves:] contains 104 quadrangular planar shapes extracted from poplar leaves in a joint collaboration with Institute for Forest Biometry and Informatics at the University of G\"ottingen, cf. \cite{Hshapes, HHM09}. \item[digits '3':] contains 30 planar shapes with 13 landmarks each, extracted from handwritten digits '3', cf. \citep[p. 318]{DM98}. \item[macaque skulls:] contains three-dimensional shapes with 7 landmarks each, of 18 macaque skulls, cf. \citep[p. 16]{DM98}. \item[iron age brooches:] contains 28 three-dimensional tetrahedral shapes of iron age brooches, cf. \citep[Section 3.5]{S96}. \end{description} \begin{figure}[h!]\centering \includegraphics[angle=-90,width=0.45\textwidth]{three_means_leaves.eps} \includegraphics[angle=-90,width=0.45\textwidth]{three_means_digit3.eps} \\ \includegraphics[angle=-90,width=0.45\textwidth]{three_means_macq.eps} \includegraphics[angle=-90,width=0.45\textwidth]{three_means_brooch.eps} \caption{\it Depicting shape means for four typical data sets: intrinsic (star), Ziezold (circle) and full Procrustes (diamond) projected to the tangent space at the intrinsic mean. The cross divides the generalized geodesic segment joining the intrinsic with the full Procrustes mean by the ratio $1:3$.\label{three_means:fig}} \end{figure} \begin{table}[h!]\centering\fbox { \footnotesize $\begin{array}{r|ccc|l} \mbox{data set} &d_{\Sigma_m^k}(\mu^{(i)},\mu^{(z)})&d_{\Sigma_m^k}(\mu^{(p)},\mu^{(z)})&d_{\Sigma_m^k}(\mu^{(p)},\mu^{(i)})&(1-\|\operatorname {\mathbb E}(Y)\|)^2\\ \hline \mbox{poplar leaves}&6.05e-05&1.83e-04&2.44e-04&5.24e-05\\ \mbox{digits '3'}&0.00154&0.00452&0.00605& 0.00155\\ \mbox{macaque skulls}&1.96e-05&5.89e-05&7.85e-05&7.59e-06\\ \mbox{iron age brooches}&0.000578&0.001713&0.002291&0.000217 \end{array}$} \caption{\it Mutual shape distances between intrinsic mean $\mu^{(i)}$, Ziezold mean $\mu^{(z)}$ and full Procrustes mean $\mu^{(p)}$ for various data sets. Last column: the concentration parameter from (\ref{reg_extr_mean_cond}), cf. also Corollary \ref{3_means_on_kendall_cor}.\label{1:3:tab}} \end{table} As clearly visible from Figure \ref{three_means:fig} and Table \ref{1:3:tab}, the approximation of Corollary \ref{3_means_on_kendall_cor} for two- and three-dimensional shapes is highly accurate for data of little dispersion (the macaque skull data) and still fairly accurate for highly dispersed data (the digits '3' data). \subsection{``Partial Blindness'' of full Procrustes and Schoenberg Means}\label{blindness:scn} In the second example we illustrate an effect of ``blindness to data'' of full Procrustes means and Schoenberg means. The former blindness is due to the affinity of the Procrustes mean to the mode in conjunction with curvature, the latter is due to non-isometry of the Schoenberg embedding. While the former effect occurs only for some highly dispersed data when the analog of condition (\ref{reg_extr_mean_cond}) is violated, the latter effect is local in nature and may occur for concentrated data as well. \begin{figure}[h!] \begin{minipage}{0.6\textwidth}\centering \includegraphics[angle=-90,width=1\textwidth]{procrustes_blind4.eps} \end{minipage} \begin{minipage}{0.05\textwidth}\hfill \end{minipage} \begin{minipage}{0.33\textwidth} \caption{\it A data set of three planar triangles (top row) with its corresponding intrinsic mean (bottom left), Ziezold mean (bottom center) and full Procrustes mean (bottom right).\label{procrustes_blind:fig}} \end{minipage} \end{figure} Reenacting the situation of Section \ref{Procrustes-convex:scn}, cf. also Example \ref{different_means_ex} and Figure \ref{different_means_fig}, the shapes of the triangles $q_1$ and $q_2$ in Figure \ref{procrustes_blind:fig} are almost maximally remote. Since the mode $q_1$ is assumed twice and $q_2$ only once, the full Procrustes mean is nearly blind to $q_2$. \begin{figure}[h!] \begin{minipage}{0.6\textwidth} \centering \includegraphics[angle=-90,width=0.3\textwidth]{schoen_sampl1.eps} \includegraphics[angle=-90,width=0.3\textwidth]{schoen_sampl2.eps} \includegraphics[angle=-90,width=0.3\textwidth]{schoen_sampl3.eps}\\ \includegraphics[angle=-90,width=1\textwidth]{schoen_blind.eps}\end{minipage} \begin{minipage}{0.05\textwidth}\hfill \end{minipage} \begin{minipage}{0.33\textwidth} \caption{\it Planar triangles $q_1= x\cos\beta - w_2\sin\beta$, $q_2= x\cos\beta + w_2\sin\beta$ and $q = x\cos\beta + w_1\sin\beta$ with $x,w_1,w_2$ from Remark \ref{Schoenberg-emb-non-iso:rm}, $\phi =0.05$, $\beta=0.3$ (top row). Intrinsic means (middle row) of sample $(q_1,q_2)$ (left) and $(q_1,q_2,q)$ (right). Schoenberg means (bottom row) of sample $(q_1,q_2)$ (left) and $(q_1,q_2,q)$ (right).\label{Schoenberg_blind:fig} } \end{minipage} \end{figure} Even though Schoenberg means have been introduced to tackle 3D shapes, the effect of ``blindness'' can be well illustrated already for 2D. To this end consider $x=x(\phi)$, $w_1=w_1(\phi)$ and $w_2=w_2(\phi)$ as introduced in Remark \ref{Schoenberg-emb-non-iso:rm}. Along the horizontal geodesic through $x$ with initial velocity $w_2$ we pick two points $q_1= x\cos\beta +w_2\sin\beta$ and $q_2= x\cos\beta - w_2\sin\beta$. On the orthogonal horizontal geodesic through $x$ with initial velocity $w_1$ pick $q = x\cos\beta' + w_1\sin\beta'$. Recall from Remark \ref{Schoenberg-emb-non-iso:rm}, that along that geodesic the derivative of the Schoenberg embedding can be made arbitrarily small for $\phi$ near $0$. Indeed, Figure \ref{Schoenberg_blind:fig} illustrates that in contrast to the intrinsic mean, the Schoenberg mean is ``blind'' to the strong collinearity of $q_1$ and $q_2$. \subsection{Discrimination Power}\label{discrimination:scn} In the ultimate example we illustrate the consequences of the choice of tangent space coordinates and the effect of the tendency of the Schoenberg mean to increase dimension by a classification simulation. To this end we apply a Hotelling $T^2$-test to discriminate the shapes of 10 noisy samples of regular unit cubes from the shapes of 10 noisy samples of pyramids with top section chopped off, each with 8 landmarks, given by the following configuration matrix $$\left(\begin{array}{cccccccc} 0& 1 &\frac{1+\epsilon}{2}& \frac{1-\epsilon}{2}& 0& 1&\frac{1+\epsilon}{2}& \frac{1-\epsilon}{2}\\ 0& 0& \frac{1-\epsilon}{2}&\frac{1-\epsilon}{2} & 1& 1& \frac{1+\epsilon}{2}&\frac{1+\epsilon}{2}\\ 0& 0& \epsilon& \epsilon& 0& 0&\epsilon&\epsilon\end{array} \right)\,$$ (cf. Figure \ref{cube_pyramid:fig}) determined by $\epsilon >0$. In the simulation, independent Gaussian noise is added to each landmark measurement. Table \ref{sim-tab} gives the percentages of correct classifications. \begin{figure}[h!] \centering \includegraphics[width=0.45\textwidth]{cube.eps} \includegraphics[width=0.45\textwidth]{pyramid.eps} \caption{\it cube (left) and pyramid of varying height $\epsilon$ (right) for classification.\label{cube_pyramid:fig}} \end{figure} \begin{table}[h!]\centering\fbox $\begin{array}{c|cccc} &\multicolumn{2}{c}{\mbox{intrinsic mean with}}&\mbox{Ziezold}&\mbox{Schoenberg}\\ \epsilon&\mbox{intrinsic}&\mbox{residual}&\mbox{mean}&\mbox{mean}\\ &\multicolumn{2}{c}{\mbox{tangent space coordinates}}\\ \hline 0.0 &70\,\% & 74\,\% & 74\,\% & 64\,\%\\ 0.2 &56\,\% & 58\,\% & 57\,\% & 51\,\%\\ 0.3 &41\,\% & 42\,\% & 42\,\% & 42\,\% \end{array}$} \caption{\it Percentage of correct classifications within 1,000 simulations each of $10$ unit-cubes and $10$ pyramids determined by $\epsilon$ (which gives the height), where each landmark is independently corrupted by Gaussian noise of variance $\sigma^2=0.2$ via a Hotelling T$^2$-test for equality of means to the significance level $0.05$.\label{sim-tab}} \end{table} As visible from Table \ref{sim-tab}, discriminating flattened pyramids ($\epsilon = 0$) from cubes ($\epsilon=1$) is achieved much better by employing intrinsic or Ziezold means rather than Schoenberg means. This finding is in concord with Theorem \ref{dim-Schoenberg-mean:thm}: samples of size $10$ of two-dimensional configurations yield Euclidean means a.s. in ${\cal P}^{7}$ which are projected to ${\cal P}^{3}$ to obtain Schoenberg means in $\Sigma_3^8$. In consequence, Schoenberg means of noisy nearly two-dimensional pyramids are essentially three dimensional. With increased height of the pyramid, ($\epsilon>0$, i.e. for more pronounced third dimension and increased proximity to the unit cube) this effect waynes and all means perform equally well (or bad). Moreover in any case, intrinsic means with intrinsic tangent space coordinates qualify less for shape discrimination than intrinsic means with residual tangent space coordinates, cf. Remark \ref{use_is_ext_not_intr_coord_rm}. The latter (intrinsic means with residual tangent space coordinates) are better or equally well behaved as Ziezold means (which naturally use residual tangent space coordinates). \begin{table}[h!]\centering \fbox{ $\begin{array}{ccc} \mbox{intrinsic mean}&\mbox{Ziezold mean} &\mbox{Schoenberg mean}\\\hline 0.24& 0.18& 0.04\end{array}$} \caption{\it Average time in seconds for the computation of means in $\Sigma_3^8$ of sample size $20$ on a PC with a $800$ MHZ CPU based on $1,000$ repetitions.\label{comp_time:tab}} \end{table} In conclusion we record the time for the computations of means in Table \ref{comp_time:tab}. While Ziezold means compute in approximately $3/4$ of the computational time for intrinsic means, Schoenberg means are obtained approximately $6$ times faster. \section{Discussion}\label{discussion-sec} By establishing stability results for intrinsic and Ziezold means on the manifold part of a shape space, a gap in asymptotic theory for general non-manifold shape spaces could be closed, now allowing for multi-sample tests of equality of intrinsic means and Ziezold means. A similar stability assertion in general is false for Procrustean means for low concentration. There is reason to believe, however, that it would be true for higher concentration. Note that the argument applied to intrinsic and Ziezold means fails for Procrustean means, since in contrast to the equations in Theorem \ref{pop_int_mean_bottom_int_mean_top:Th} the sum of Procrustes residuals is in general non-zero. Loosely speaking, the findings on dimensionality condense to {\it \begin{itemize} \item[--] Procrustean means may decrease dimension by 2 or more, \item[--] intrinsic and Ziezold means decrease dimension at most by 1, in particular, they preserve regularity, \item[--] Schoenberg means tend to increase up to the maximal dimension possible. \end{itemize}} Due to the proximity of Ziezold and intrinsic means on Kendall's shape spaces in most practical applications, taking into account that the former are computationally easier accessable (optimally positioning and Euclidean averaging in every iteration step) than intrinsic means (optimally positioning and weighted averaging in every iteration step), Ziezold means can be preferred over intrinsic means. They may be even more preferred over intrinsic means, since Ziezold means naturally come with residual tangent space coordinates which may allow in case of intrinsic means for a higher finite power of tests than intrinsic tangent space coordinates. Computationally much faster (not relying on iteration at all) are Schoenberg means which are available for Kendall's reflection shape spaces. As a drawback, however, Schoenberg means, seem less sensitive for dimensionality of configurations considered than intrinsic or Ziezold means. In particular for problems involving small sample sizes $n$ and a large number of parameters $p$ as currently of high interest in statistcal applications, involving (nearly) degenerate data, Ziezold means may also be preferred over Schoenberg means due to higher power of tests. Finally, note that Ziezold means may be defined for the shape spaces of planar curves introduced by \cite{ZR72}, which are currently of interest e.g. \cite{KSMJ04} or \cite{SCC06}. Employing Ziezold means there, a computational advantage greater than found here can be expected since the computation of iterates of intrinsic means involves computations of geodesics which themselves can only be found iteratively. \section*{Acknowledgment} The author would like to thank Alexander Lytchak for helpful advice on differential geometric issues. \section{Introduction}\label{intro-scn} The analysis of shape may be counted among the very early activities of mankind; be it for representation on cultural artefacts, or for morphological, biological and medical application, to name only two of its driving forces. In modern days shape analysis is gaining increased momentum in computer vision, image analysis, biomedicine and many other fields. For a fairly recent overview over key aspects cf. \cite{KY06}. To begin with, already the concept of an expected, averaged or \emph{mean} shape is surprisingly non trivial and not at all canonical. This is due to the fact that spaces of shapes usually admit several natural structures, and, in many cases none of them is that of a Euclidean space. While some structures are Euclidean (e.g. \cite{M70}, \cite{HKD02}, \cite{HHGMS07}), structures that may be considered benign are that of Riemanian manifolds (e.g. \cite{BN78}, \cite{K84}). Many shape space, however, admit only the structure of a Riemannian manifold -- the \emph{pre-shape space} of configurations -- modulo the isometric action of a Lie group, conveying shape equivalence. We call the quotient a \emph{shape space}. It carries a canonical quotient structure of a union of manifold strata of different dimensions, which give in general a Riemannian manifold -- the \emph{manifold part} -- with singularities at some of which the curvature may tend to infinity. Prominent example are Kendall's three- and higher-dimensional shape spaces, cf. \citet{KBCL99}, and the spaces of closed planar curves, see \citet{ZR72}. In a Euclidean space, there is a clear and unique concept of a mean in terms of least squares minimization: the arithmetic avarage. Generalizing to manifolds, however, with every embedding in a Euclidean space come specific concepts of \emph{extrinsic means} and \emph{residual means}, and with every manifold structure a specific concept of an \emph{intrinsic mean}. Carrying statistics over to manifolds, strong consistency and central limit theorems (CLTs) for extrinsic, residual and intrinsic means have been derived by \cite{J88}, \cite{HL96,HL98} as well as by \cite{BP03,BP05}. Under quotienting these means generalize to \emph{Ziezold means} and \emph{Procrustes means}, respectively, and again, to \emph{intrinsic means}. For a CLT to hold, obviously, a manifold structure is inevitable. Due to strong consistency, which has been established by \cite{Z77} on general quasi-metrical spaces, for a \emph{one-sample test} for a specific mean shape on the manifold part, it may be assumed that sample means eventually lie on the manifold part as well, thus making the above CLT available. Multi-sample tests, however, could not be theoretically justified because it remained unclear whether means of random shapes on the manifold part come to lie on the manifold part again. As the key result in this paper it is shown that intrinsic and Ziezold means lie on the manifold part whenever the manifold part is assumed with non-zero probability. In consequence the CLT for manifolds can be applied in general, thus making multi-sample tests for these mean shapes possible. We note that specifically for the non-manifold Kendall reflection shape spaces, \emph{Schoenberg means} as have been recently introduced by \cite{BandPat05} as well as by \cite{DKLW08} also allow for asymptotic inference. This paper is structured as follows. In motivation in Section \ref{Kendalls-ss:scn}, Kendall's shape spaces are introduced along with the specific application of the key result. The following Section \ref{Frechet-rho-means:scn} lays out the various concepts of means in a generic shape space setting. In Section \ref{convex:scn} the stability theorem, namely that \begin{center} intrinsic means preserve dimension \end{center} is established. The proof is based on lifting a distribution on the shape space to the pre-shape space and subsequently exploiting the fact that intrinsic means are zeroes of an integral involving the Riemann exponential. The similar argument can be applied to Ziezold means but not to Procrustes means. The population case is more intrigued, however, as the lifting requires the concept of tubular neighborhoods admitting slices. In Section \ref{ext-means:scn}, extrinsic means are discussed, in particular that Schoenberg means are due to a non-isometric embedding and that they tend to ``increase the dimension''. Section \ref{local:scn} tackles local effects of curvature: Section \ref{tangent:scn} motivates why residual tangent space coordinates may yield a higher finite power than intrinsic tangent space coordinates. The closeup on spherical means in Section \ref{1:3:scn} reveals that, if the following means are unique, the generalized geodesic segment between the intrinsic mean and the Procrustes mean (closer to the data) is in approximation divided by the Ziezold mean by the ration $1:3$. Section \ref{class_data_simulations_scn} illustrates the practical effects of the theoretical results using classical data-sets as well as simulations. \section{Motivation: Kendall's Shape Spaces}\label{Kendalls-ss:scn} In the statistical analysis of similarity shapes based on landmark configurations, geometrical $m$-dimensional objects (usually $m=2,3$) are studied by placing $k>m$ \emph{landmarks} at specific locations of each object. Each object is then described by a matrix in the space $M(m,k)$ of $m\times k$ matrices, each of the $k$ columns denoting an $m$-dimensional landmark vector. $\langle x,y\rangle := \operatorname {tr}(xy^T)$ denotes the usual inner product with norm $\|x\| = \sqrt{\langle x,x\rangle}$. For convenience and without loss of generality for the considerations below, only \emph{centered} configurations are considered. Centering can be achieved by multiplying with a sub-Helmert matrix ${\cal H}\in M(k,k-1)$ from the right, yielding a configuration $x{\cal H}$ in $M(m,k-1)$. For this and other centering methods cf. \citet[Chapter 2]{DM98}. Excluding also all configurations with all landmarks coinciding gives the space of \emph{configurations} \begin{eqnarray* F_m^k&:=& M(m,k-1) \setminus \{0\} \,. \end{eqnarray*} Since only the similarity shape is of concern, we may assume that all configurations are contained in the unit sphere $ S_m^k :=\{x\in M(m,k-1): \|x\|=1\}$. Then, \emph{Kendall's shape space} is the canonical quotient $$\Sigma_m^k := S_m^k/SO(m) = \{[x]:x\in S_m^k\}\mbox{ with the \emph{fiber} } [x] = \{gx:g\in SO(m)\}\,.$$ In some applications reflections are also filtered out giving \emph{Kendall's reflection shape space} $$R\Sigma_m^k := \Sigma_m^k/\{e,\widetilde{e}\}= S_m^k/O(m)\,.$$ Here, $O(m) = \{g\in M(m,m): g^Tg=e\}$ denotes the orthogonal group with the unit matrix $e=\operatorname {diag}(1,\ldots,1)$, $\widetilde{e} = \operatorname {diag}(-1,1,\ldots,1)$ and $SO(m) = \{g\in O(m): \det(g)=1\} $ is the special orthogonal group. For $1\leq j \leq m < k$ consider the (non-unique) embedding \begin{eqnarray}\label{pre-shape-emb:eq} \begin{array}{rclcrcl} S_j^k &\hookrightarrow& S_m^k&:& x&\mapsto&\left(\begin{array}{c}x\\\hline 0\end{array}\right)\end{array}\, \end{eqnarray} giving rise to a unique and canonical embedding of $\Sigma_j^k\hookrightarrow \Sigma_m^k$ which is isometric w.r.t. the canonical intrinsic distance, the Procrustes distance and the Ziezold distance defined in Section \ref{Frechet-rho-means:scn}, cf. \citet[pp. 29, 206]{KBCL99}. In consequence, intrinsic means, Procrustes means and Ziezold means (also defined in Section \ref{Frechet-rho-means:scn}) of random elements taking values in $\Sigma_j^k\subset \Sigma_m^k$, are also contained within $\Sigma_j^k$: \begin{Rm}\label{decr_dim:rm} Suppose that $[X]$ is a random shape on $\Sigma_m^k$ assuming shapes of configurations of dimension less than or equal to $j$ ($1\leq j \leq m$) with probability one, then every intrinsic, Procrustes and Ziezold mean shape of $[X]$ corresponds to a configuration of dimension less than or equal to $j$. \end{Rm} The main result of this paper in form Corollaries \ref{sample_mean_reg:cor} and \ref{pop_mean_reg:cor} applied to Kendall's spaces states that intrinsic and Ziezold means are also non decreasing in dimension. In view of Ziezold means, Remark \ref{kendall_book:rm} provides invariant optimal positioning and (\ref{Ziez-inj:cond}) on page \pageref{Ziez-inj:cond} is valid on the pre-shape sphere. The technical definition of a cut locus on the quotient and a tubular neighborhood admitting a slice can be found in Section \ref{prelims:scn}. \begin{Th}[Intrinsic and Ziezold Kendall Means Preserve Dimension]\label{Kendall_mean_dim} Suppose that $[X]$ is a random pre-shape on $\Sigma_m^k$, $m>k$ with intrinsic or Ziezold mean shape $[\mu] \in \Sigma_m^k$, $\mu \in S_m^k$. Moreover suppose that $[X]$ either assumes discrete values outside the cut locus of $[\mu]$ a.s. or that it is supported by a projection of a tubular neighborhood around $\mu$ admitting a slice. If $[X]$ assumes shapes corresponding to non-degenerate configurations up to dimenson $j$ ($1\leq j\leq m$) with non-zero probability then $\mu$ corresponds to a non-degenerate $j$-dimensional configuration. \end{Th} \section{Fr\'echet $\rho$-means and General Shape Spaces}\label{Frechet-rho-means:scn} In the previous section we introduced Kendall's \emph{shape} and \emph{reflection shape space} based on invariance under similarity transformations; e.g. invariance under congruence transformations only leads to Kendall's \emph{size-and-shape space}. More generally in image analysis, invariance may also be considered under the affine or projective group, cf. \cite{MP01,MP05,MPPPR07}. A different yet also very popular popular set of shape spaces for two-dimensional configurations modulo the group of similarites has been introduced by \cite{ZR72}. Instead of building on a finite dimensional Euclidean matrix space modeling landmarks, the basic ingredient of these spaces modeling closed planar unit speed curves is the infinite dimensional Hilbert space of Fourier series. In practice for numerical computation, only finitely many Fourier coefficients are considered. To start with, a shape space is a metric space $(Q,d)$. For this entire paper suppose that $X, X_1,X_2,\ldots$ are i.i.d. random elements mapping from an abstract probability space $(\Omega,\cal A,P)$ to $(Q,d)$ equipped with its self understood Borel $\sigma$-field. Moreover, denote by $\mathbb E(Y)$ the classical expected value of a random element $Y$ on a $D$-dimensional Euclidean space $\mathbb R^D$, if existent. \begin{Def}\label{Frechet_means:def} For a continuous function $\rho:Q\times Q \to [0,\infty)$ define the \emph{set of population Fr\'echet $\rho$-means} by $$ E^{(\rho)}(X) = \operatorname {argmin}_{\mu\in Q} \mathbb E\big(\rho(X,\mu)^2\big) \,.$$ For $\omega\in \Omega$ denote by $$ E^{(\rho)}_n(\omega) = \operatorname {argmin}_{\mu\in Q} \sum_{j=1}^n \rho\big(X_j(\omega),\mu\big)^2$$ the \emph{set of sample Fr\'echet $\rho$-means}. \end{Def} By continuity of $\rho$, the mean sets are closed random sets. For our purpose here, we rely on the definition of \emph{random closed sets} as introduced and studied by \cite{Choq54}, \cite{Kend74} and \cite{Math75}. Since their original definition for $\rho=d$ by \cite{F48} such means have found much interest. \paragraph{Intrinsic means.} Independently, for a Riemannian manifold with geodesic distance $d=\rho$, \cite{KN69} defined the corresponding means as \emph{centers of gravity} which are nowadays also well known as \emph{intrinsic means} (\cite{BP03,BP05}). \paragraph{Extrinsic means.} W.r.t. the chordal or \emph{extrinsic metric} $d=\rho$ due to an embedding of a Riemannian manifold in an ambient Euclidean space, such means have been called \emph{mean locations} by \cite{H91} or \emph{extrinsic means} by \cite{BP03}. More precisely, let $Q=M\subset \mathbb R^D$ be a Riemannian manifold embedded in a Euclidean space $\mathbb R^D$ and let $\Phi: R^D \to M$ denote the orthogonal projection, $\Phi(x) = \inf_{p\in M}\|x-p\|$. For any Riemannian manifold an embedding that is even isometric can be found for $D$ sufficiently large, see \cite{Na56}. Due to an extension of Sard's Theorem by \citet[p.12]{BP03} for a closed manifold, $\Phi$ is uni-valent up to a set of Lebesgue measure zero. Then the set of extrinsic means is given by the set of images $\Phi\big(\mathbb E(Y)\big)$ where $Y$ denotes $X$ viewed as taking values in $\mathbb R^D$ (cf. \cite{BP03}). \paragraph{Residual means.} In this context, setting $\rho(p,q) = \|d\Phi_q(p-q)\|$ ($p,q\in M$) with the derivative $d\Phi_q$ at $q$ yielding the orthogonal projection to the embedded tangent space $T_q\mathbb R^D \to T_qM\subset T_q\mathbb R^D$, call the corresponding mean sets $E^{(\rho)}(X)$ and $E^{(\rho)}_n(\omega)$, the sets of \emph{residual population means} and \emph{residual sample means}, respectively. For two-spheres, $\rho(p,q)$ has been studied under the name of \emph{crude residuals} by \cite{J88}. In general, on unit-spheres \begin{eqnarray}\label{res_dist_sphere:def} \rho(p,q) =\|p - \langle p,q\rangle q\| = \sqrt{1-\langle p,q\rangle^2} = \rho(q,p) \end{eqnarray} is a quasi-metric (symmetric, vanishing on the diagonal $p=q$ and satisfying the triangle inequality). ~\\ Due to curvature these concepts of means are essentially non-convex in the following sense. \begin{Rm}\label{spherical_means:rm} To a random point on a two-sphere, uniformly distributed along its equator, both north and south pole are intrinsic and extrinsic means. \end{Rm} Residual means, however, seem less affected. \begin{Th}\label{res_means_sphere:rm} If a random point $X$ on a unit sphere is a.s. contained in a unit subsphere $S$ then $S$ contains as well every residual mean of $X$.\end{Th} \begin{proof} Suppose that $x = v+\nu$ is a residual mean of $X$ with $v/\|v\| \in S$ and $\nu$ normal to $S$. Since $1-\langle X,v+\nu\rangle^2 = 1-\langle X,v\rangle^2 \geq 1 - \langle X,v\rangle^2/\|v\|^2$ a.s. with equality if and only if $\nu=0$, the assertion follows at once from (\ref{res_dist_sphere:def}). \end{proof} Let us now incorporate more of the structure common to shape spaces: $M$ is a complete connected $D$-dimensional Riemannian manifold with geodesic distance $d_M$ on which a Lie group $G$ acts properly and isometrically from the left. Then we call the canonical quotient $$\pi: M \to Q:= M/G = \{[p]: p \in M\}\mbox{ where } [p] = \{gp: g\in G\}\,,$$ a \emph{shape space}. As a consequence of the isometric action we have that $d_M(gp,q) = d(p,g^{-1}q)$ for all $p,q\in M$, $g\in G$. For $p,q\in M$ we say that $p$ is in \emph{optimal position} to $q$ if $d_M(p,q) =\min_{g\in G}d_M(gp,q)$, the minimum is attained in consequence of the proper action. As is well known (e.g. \citet[p. 179]{Bre72}) there is an open and dense submanifold $M^*$ of $M$ such that the canonical quotient $Q^* = M^*/G$ restricted to $M^*$ carries a natural manifold structure also being open and dense in $Q$. Elements in $M^*$ and $Q^*$, respectively, are called \emph{regular}, the complementary elements are \emph{singular}; $Q^*$ is the \emph{manifold part}. \paragraph{Intrinsic means on shape spaces.} The canonical distance $$d_Q([p],[q]) :=\min_{g\in G}d_M(gp,q) = \min_{g,h\in G}d_M(gp,hq)$$ is called \emph{intrinsic distance} and the corresponding $d_Q$-Fr\'echet mean sets are called \emph{intrinsic means}. Note that the intrinsic distance is equal to the canonical geodesic distance on $Q^*$. \paragraph{Ziezold and Procrustes means.} Now, assume that we have an embedding with orthogonal projection $\Phi: R^D \to M\subset\mathbb R^D$ as above. If the action of $G$ is isometric w.r.t. the extrinsic metric, i.e. if $\|gp-gq\| = \|p-q\|$ for all $p,q\in M$ and $g\in G$ then call \begin{eqnarray*} \rho^{(z)}_Q([p],[q])&:=&\min_{g\in G}\|gp-q\|\\ \rho^{(p)}_Q([p],[q])&:=&\min_{g\in G}\|d\Phi_q(gp-q)\| \end{eqnarray*} the \emph{Ziezold distance} and the \emph{Procrustes distance} on $Q$, respectively. The corresponding Fr\'echet mean sets are the \emph{Ziezold means} and the \emph{Procrustes means}, respectively. We say that \emph{optimal positioning is invariant} if $$d_M(g^*p,q) = \min_{g\in G}d_M(gp,q) \Leftrightarrow \|g^*p-q\| = \min_{g\in G}\|gp-q\|$$ for all $p,q \in M$ and $g^*\in G$. \begin{Rm}\label{kendall_book:rm} Indeed for $Q=\Sigma_m^k$, optimal positioning is invariant (cf. \citet[p. 206]{KBCL99}), Procrustes means coincide with classical Procrustes means introduced by \cite{Gow} and Ziezold means coincide with means as introduced by \cite{Z94}. Moreover for $Q=\Sigma_2^k$, Procrustes means agree with extrinsic means w.r.t. the Veronese-Whitney embedding, cf. \cite{BP03} and Section \ref{ext-means:scn}. \end{Rm} \section{Stability Results for Intrinsic Means}\label{convex:scn} In this section we derive the stability results underlying Theorem \ref{Kendall_mean_dim}. To this end we need to recall how a non-manifold shape space is made up from manifold strata of varying dimensions. \subsection{Preliminaries}\label{prelims:scn} As before, let $M$ be a complete connected Riemannian manifold called the \emph{pre-shape space} with intrinsic metric $d_M$ on which a Lie group $G$ acts properly and isometrically giving rise to the canonical quotient $Q=M/G$ called the \emph{shape space} with intrinsic metric $d_Q$. The canonical projection is denoted by $\pi :M\to Q$. $T_pM$ is the tangent space of $M$ at $p\in M$ and $\exp_p$ denotes the \emph{Riemannian exponential} at $p$. On a complete and connected manifold $M$ we have for every $p'\in M$ that there is $v' \in T_pM$ such that $p'= \exp_p v'$. $v' = \exp^{-1}_pp'$ of minimal modulus is uniquely determined as long as $p\in M \setminus C(p)$ with the \emph{cut locus} $C(p)$ of $p$ which is a subset of measure zero in $M$. E.g. on a sphere, the cut locus of any point is its antipode. For $q\in Q$ let $$C(q) := \{[p']: p'\in C(p) \mbox{ is in optimal position to some } p\in q\}\,$$ be the \emph{cut locus on the quotient} of $q\in Q$. Next we recollect consequences of the Lie group action, see \cite{Bre72}. The stability result in the sample case (below in Section \ref{sample_convex:scn}) rests on (A) and (B) alone. \begin{enumerate} \item[(A)] With the \emph{isotropy group} $I_p=\{g\in G: gp =p\}$ for $p\in M$, every orbit carries the natural structure of a coset space $[p] \cong G/I_p$. Moreover, $p' \in M$ is of \emph{orbit type} $(G/I_p)$ if $I_{p'} = gI_pg^{-1}=I_{gp}$ for a suitable $g\in G$. If $I_p \subset I_{gp'}$ for suitable $g\in G$ then $p'$ is of \emph{lower orbit type} than $p$. \item[(B)] The pre-shapes of equal orbit type $$M^{(I_p)} := \{p'\in M:\mbox{ $p'$ is of orbit type }(G/I_p)\}$$ and the corresponding shapes $Q^{(I_p)}:=\{[p']:p'\in M^{(I_p)}\}$ are manifolds in $M$ and $Q$, respectively. \item[(C)]The orthogonal complement $H_pM$ in $T_pM$ of the tangent space $T_p[p]$ along the orbit is called the \emph{horizontal space}: $T_pM = T_p[p]\oplus H_pM$. \item[(D)] The \emph{Slice Theorem} (cf. \cite{P60}) states that every $p\in M$ has a \emph{tubular neighborhood} $[p]\subset U \subset M$ such that with a suitable open subset $D\subset H_pM$ the \emph{twisted product} $\exp_p D\times_{I_p} [p]$ is diffeomorphic with $U$. Here, the twisted product is the natural topological quotient of the product space $\exp_p D\times [p]$ modulo the equivalence $$ (\exp_pv,gp) \sim_{I_p} (\exp_pv', g'p)~\Leftrightarrow \exists h \in I_p\mbox{ such that } v'=dh v,~g'=gh^{-1}\,.$$ For short we say that the tubular neighborhood $U$ \emph{admits a slice} \mbox{$\exp_p D$} \emph{through $p$}. \item[(E)] If a tubular neighborhood $U$ of $p$ admits a slice $\exp_p D$ then every $p' \in \exp_pD$ is in optimal position to $p$. Moreover, all points $p'\in U$ are of orbit type larger than or equal to the orbit type of $p$ and only finitely many orbit types occur in $U$. If $p$ is regular, i.e. of maximal orbit type, then the quotient is trivial: $\exp_p D \times_{I_p} [p] =\exp_p D \times [p]$. \end{enumerate} Finally, we link intrinsic to Ziezold means. The gradient of the mapping $f_{int} : T_pM \to [0,\infty)$ defined by $f^{(p)}_{int}(v) = d_M(p,\exp_pv)^2$ is given by $\operatorname {grad} f^{(p)}_{int}(v) = 2v$. Hence, we have for $p_1,p_2 \in M\setminus C(p)$ that $$ \operatorname {grad} f^{(p)}_{int}(\exp_p^{-1}p_1) = \operatorname {grad}(f^{(p)}_{int})(\exp_p^{-1}p_2) ~\Leftrightarrow~ p_1=p_2$$ In view of Ziezold means let $f^{(p)}_{ext} : T_pM \to [0,\infty)$ be defined by $f^{(p)}_{ext}(v)=\|p-\exp_pv\|^2/2$. Mimicking the above property introduce the following condition \begin{eqnarray}\label{Ziez-inj:cond} \operatorname {grad}(f^{(p)}_{ext})(\exp_p^{-1}p_1) = \operatorname {grad}(f^{(p)}_{ext})(\exp_p^{-1}p_2) &\Leftrightarrow& p_1=p_2 \end{eqnarray} for $p_1,p_2 \in M\setminus C(p)$. E.g. on spheres, (\ref{Ziez-inj:cond}) is valid. Using the residual distance instead, the analog to (\ref{Ziez-inj:cond}) is only true on half-spheres. \subsection{Stability of Sample Means}\label{sample_convex:scn} For short, for sampled points $p_1,\ldots, p_n \in M$ denote the set of intrinsic or extrinsic sample means, respectively, by $E(p_1,\ldots,p_n)$; similarly for $q_1,\ldots, q_n \in Q$ denote the set of intrinsic or Ziezold sample means, respectively, by $E(q_1,\ldots,q_n)$. Obviously, both sets are non-void. \begin{Lem}\label{sample_int_mean_bottom_int_mean_top:lem} In case of extrinsic and Ziezold means assume that optimal positioning is invariant. Let $q_1\ldots, q_n \in Q$ and $\overline{q}\in E(q_1,\ldots,q_n)$. If $\overline{p} \in \overline{q}$ and $p_1\in q_1,\ldots, p_n\in q_n$ are in optimal position to $\overline{p}$ then $\overline{p} \in E(p_1,\ldots,p_n)$. \end{Lem} \begin{proof} The proof in case of extrinsic and Ziezold means is completely analogous to the following case of intrinsic means. If the assertion would not be true, we would have with $p \in E(p_1,\ldots,p_n)$ that \begin{eqnarray*} \sum_{j=1}^nd_Q([p], q_j)^2 &=& \sum_{j=1}^nd_M(p,g_jp_j)^2\\ &<&\sum_{j=1}^n d_M(\overline{p},p_j)^2 ~=~ \sum_{j=1}^nd_Q(\overline{q}, q_j)^2 \end{eqnarray*} for $g_1p_1,\ldots, g_np_n$ in optimal position to $p$, a contradiction to the hypothesis $\overline{q}\in E(q_1,\ldots,q_n)$ \end{proof} The following necessary condition is taken from \cite[p. 110]{KN69}. \begin{Lem}\label{finite-kn69:lem} In case of intrinsic means let $\overline{p} \in M$ and $p_1,\ldots,p_n \in M\setminus C(\overline{p})$ with $\overline{p} \in E(p_1,\ldots,p_n)$ for a complete and connected manifold $M$. Then $$\sum_{j=1}^n \exp_{\overline{p}}^{-1}p_j =0\,.$$ \end{Lem} For extrinsic means the analog condition is \begin{eqnarray}\label{sum-extr-mean:eq} \sum_{j=1}^n f^{(\overline{p})}_{ext}(\overline{p},\exp_{\overline{p}}^{-1}p_j) &=&0\,. \end{eqnarray} \begin{Th}\label{sample_mean_iso_grp:thm} Suppose that $G$ is a Lie group acting properly and isometrically on a complete connected Riemannian manifold $M$ giving rise to the natural quotient $Q=M/G$. If $\overline{q}\in E(q_1,\ldots,q_n)$ is an intrinsic sample mean of $q_1,\ldots, q_n \in Q\setminus C(\overline{q})$ then \begin{eqnarray*}I_{\overline{p}}\subset \cap_{j=1}^n I_{p_j}\,\end{eqnarray*} with arbitrary $\overline{p}\in \overline{q}$ and $p_1\in q_1,\ldots,p_n\in q_n$ in optimal position to $\overline{p}$. In case of invariant optimal positioning and under condition (\ref{Ziez-inj:cond}) the same assertion is true if $\overline{q}$ is a Ziezold sample mean. \end{Th} \begin{proof} The proof in case of Ziezold means is completely analogous to the following case of intrinsic means, using (\ref{sum-extr-mean:eq}) instead of Lemma \ref{finite-kn69:lem}. If the assertion would be false, we may assume that $I_{\overline{p}} \not\subset I_{p_1}$, i.e. that there is $g\in G$ with $g\overline{p} =\overline{p}$ but $gp_1\neq p_1$. In consequence of Lemma \ref{sample_int_mean_bottom_int_mean_top:lem}, $\overline{p}$ is an intrinsic mean of both $p_1,\ldots, p_n$ and $gp_1, p_2,\ldots,p_n$ , both being in optimal position to $\overline{p}$. Since by hypothesis, $\exp^{-1}_{\overline{p}}(gp_1) \neq \exp^{-1}_{\overline{p}}(p_1)$ is well defined, we have in consequence of Lemma \ref{finite-kn69:lem} the contradiction \begin{eqnarray*} 0 &=& \exp^{-1}_{\overline{p}}gp_1+ \sum_{j=2}^n \exp_{\overline{p}}p_j\\ &=& \exp^{-1}_{\overline{p}}gp_1 - \exp^{-1}_{\overline{p}}p_1 + \sum_{j=1}^n \exp_{\overline{p}}p_j ~\neq~0\,. \end{eqnarray*} \end{proof} \begin{Cor}\label{sample_mean_reg:cor} Suppose that $G$ is a Lie group acting properly and isometrically on a complete connected Riemannian manifold $M$ giving rise to the natural quotient $Q=M/G$. Then every intrinsic mean $\overline{q}\in Q$ of a sample $q_1,\ldots,q_n \in Q\setminus C(\overline{q})$ is regular if the sample contains at least one regular point. In case of invariant optimal positioning and under condition (\ref{Ziez-inj:cond}) the same assertion is true for Ziezold sample means. \end{Cor} \begin{proof} The assertion is a consequence of Theorem \ref{sample_mean_iso_grp:thm} and the well known fact that regular points $p^*\in M^*$ are characterized by the property that for every $p\in M$ there is a $g=g_p \in G$ such that $gI_{p^*}g^{-1} \subset I_p$: if additionally $p$ is in optimal position to $p^*$ then $g\in I_p$. \end{proof} \subsection{Stability of Population Means} \begin{Th}\label{population_mean_iso_grp:thm} Assume that $X$ is a random element on the canonical quotient $Q=M/G$ due to the isometric and proper action of a Lie group $G$ on a complete Riemannian manifold $M$. Let $p\in M$ and suppose that $X$ is supported on $\pi(U)$ with a tubular neigborhood $U\in M$ of $[p]$ that admits a slice $\exp_p D \times_{I_p} [p] \cong U$, $D\in H_pM$. If ${\cal P} \{X\in Q^{(I_{p'})}\} \neq 0$ for some $p'\in U$ and if either $[p]$ \begin{enumerate} \item[(i)] is an intrinsic mean of $X$ or \item[(ii)] is a Ziezold mean of $X$ while optimal positioning is invariant and (\ref{Ziez-inj:cond}) is valid, \end{enumerate} then $p'$ is of lower orbit type than $p$. \end{Th} The proof of Theorem \ref{population_mean_iso_grp:thm} further below relies on the two following lemmas for which we first develop an additional concept. In the following, \emph{measurable} will refer to the corresponding Borel $\sigma$-algebras, respectively. \begin{Def}\label{hor-lift:def} Call a measurable subset $L \subset M$ a \emph{measurable horizontal lift} of a subset $R$ of $M/G$ \emph{in optimal position to $p \in M$} if \begin{enumerate} \item the canonical projection $L \to R\subset M/G$ surjective, \item every $p' \in L$ is in optimal position to $p$, \item every orbit $[p']$ of $p'\in L$ meets $L$ once. \end{enumerate} \end{Def} \begin{Lem}\label{hor_measurable_lift:lem} Let $U\subset M$ be a tubular neighborhood about $p\in M$ that admits a slice $\exp_p D \times_{I_p} [p] \cong U$. Then, there is a measurable horizontal lift $L\subset \exp_pD$ of $\pi(U)$. \end{Lem} \begin{proof} If $p$ is regular, then $L=\exp_pD $ has the desired properties, cf. (E) above. Now assume that $p$ is not of maximal orbit type. W.l.o.g. assume that $D$ contains the closed ball $B$ of radius $r>0$ with bounding sphere $S = \partial B$ and that there are $p_1,\ldots, p_J \in \exp_p(S)$ having the distinct orbit types orccuring in $S$. $S^{j}$ denotes all points on $S$ of orbit type $(G/I_{p_j})$, $j=1,\ldots,J$, respectively. Observe that each $S^{j}$ is a manifold on which $I_p$ acts properly and isometrically. Hence for every $1\leq j\leq J$, there is a finite ($K_j<\infty$) or countable ($K_j=\infty$) sequence of tubular neighborhoods $U^j_k\subset S^{j}$ covering $S^{j}$, admitting trivial slices $$\exp^{S^j}_{p^j_k}D^j_k \times \{gp^j_k:g\in I_{p_k^j}\} \cong U^j_k, ~~1\leq k \leq K_j\,.$$ Here, $\exp^{S^j}_{p^j_k}$ denotes the Riemann exponential of $S^j$. Defining a disjoint sequence \begin{eqnarray*} \widetilde{U}^{j}_1&:=& U_1^j\\ \widetilde{U}^{j}_{k+1}&:=& U_{k+1}^j\setminus \widetilde{U}^{j}_{k}\mbox{ for $1\leq k\leq K_j-1 $} \end{eqnarray*} exhausting $S^j$ we obtain a corresponding sequence of disjoint measurable sets $\widetilde{D}^{j}_{k}$ with $$\exp^{S^j}_{p^j_k}\widetilde{D}^j_k \times \{gp^j_k:g\in I_{p^j_k}\} \cong \widetilde{U}^j_k, ~~1\leq k \leq K_j\,.$$ Setting $$L^{j}_{k} := \exp^{S^j}_{p^j_k}\widetilde{D}^j_k\mbox{ and } L^j := \bigcup_{k=1}^{K_j}L_k^j$$ observe that every $p' \in S^j$ has a unique lift in $L^j$ which is contained in a unique $L^{j}_{k}$. This lift is by construction in optimal position to $p$. If $gp' \in L^{j}_{k'}$ for some $g\in G$ and $1\leq k'\leq K_j$ we have by the disjoint construction of $\widetilde{U}^{j}_{k}$ and $\widetilde{U}^{j}_{k'}$ that $k=k'$, hence the isotropy groups of $gp'$ and $p'$ agree, yielding $gp'=p'$. In consequence, $L^j$ is a measurable horizontal lift of $S^j$ in optimal position to $p$. Since every horizontal geodesic segment $t\mapsto \exp_p(tv)$, $v\in H_pM$ contained in $\exp_pD$ features a constant isotropy group, except possibly for the initial point we obtain with the definition of $$L :=\{\exp_p(tv)\in \exp_pD: v\in \exp_p^{-1}(L^j) \mbox{ for some $1\leq j \leq J$}, t\geq 0\}$$ a measurable horizontal lift of $\pi(U)$ in optimal position to $p$. \end{proof} The following Lemma \ref{pop_int_mean_bottom_int_mean_top:lem} is the analog in the population case of Lemmas \ref{sample_int_mean_bottom_int_mean_top:lem} and \ref{finite-kn69:lem}. The proof goes similar. \begin{Lem}\label{pop_int_mean_bottom_int_mean_top:lem} Assume that $X$ is a random element on the canonical quotient $Q=M/G$ due to the isometric and proper action of a Lie group $G$ on a complete Riemannian manifold $M$. Moreover, assume that there is $p\in M$ with tubular neigborhood $U\in M$ of $[p]$ that admits a slice $\exp_p D \times_{I_p} [p] \cong U$ such that $X$ is supported by $\pi(U)$. With a measurable horizontal lift $L$ of $\pi(U)$ define the random element $Y$ on $L\subset M$ by $\pi\circ Y = X$. If $[p]$ is an intrinsic mean of $X$ on $Q$, then $p$ is an intrinsic mean of $Y$ on $M$ and $$\mathbb E(\exp^{-1}_pY)=0\,.$$ If $[p]$ is a Ziezold mean of $X$ on $Q$ and optimal positioning is invariant, then $p$ is an extrinsic mean of $Y$ on $M$ and $$\mathbb E\big(\operatorname {grad}(f_{ext})(\exp^{-1}_pY)\big)=0\,.$$ \end{Lem} \noindent {\it Proof of Theorem \ref{population_mean_iso_grp:thm}.} The proof in case of Ziezold means is completely analogous to the following case of intrinsic means. With the hypotheses of Theorem \ref{population_mean_iso_grp:thm}, suppose that $L$ is a horizontal measurable lift of $\pi(U)$ guaranteed by Lemma \ref{hor_measurable_lift:lem} through an intrinsic mean $p\in M$ of the random element $Y$ on $M$ defined as in Lemma \ref{pop_int_mean_bottom_int_mean_top:lem} with $[p]\in E^{d_Q}(X)$. With the notation of Lemma \ref{hor_measurable_lift:lem} and its proof, if the assertion of the Theorem would be false, w.l.o.g. there would be a $q_{j} \in S^{j}$ with $gp_{j} \neq p_{j} $ for some $g\in I_p$ and $P\{Y\in M^{j}\}>0$ for some $1\leq {j}\leq J$ with $$M^j :=\{\exp_p(tv)\in \exp_p(D): v\in \exp_p^{-1}(L^j), t\geq 0\}$$ In particular, w.l.o.g. in the proof Lemma \ref{hor_measurable_lift:lem}, we may choose an arbitrarily small $L_{j}^{k}$ around $p_{j}$ such that $$ \int_{M^{j}_{k}} \big(\exp^{-1}_pY - \exp^{-1}_p (gY)\big) \,dP_Y~\neq~ 0\,.$$ with $M_k^j :=\{\exp_p(tv)\in \exp_p(D): v\in \exp_p^{-1}(L^j_k), t\geq 0\}$. Suppose that $L$ is obtained as in the proof of Lemma \ref{hor_measurable_lift:lem} by using $L^{j}_{k}$ and suppose that $L'$ is obtained from $L$ by replacing $L^{j}_{k}$ with $gL^{j}_{k}$. Then $L'$ is also a measurable horizontal lift in optimal position to $p$. Moreover with the lifts $Y'$ of $X$ to $L'$ and $Y$ of $X$ to $L$ guaranteed by Lemma \ref{pop_int_mean_bottom_int_mean_top:lem}, we have that \begin{eqnarray*} 0&=&\int_{L} \exp^{-1}_pY\,dP_Y - \int_{L'} \exp^{-1}_pY'\,dP_Y'\\ &=&\int_{M^{j}_{k}} \big(\exp^{-1}_pY - \exp^{-1}_p (dgY)\big) \,dP_Y~\neq~ 0\,, \end{eqnarray*} a contradiction to the above, completing the proof. \qed We have at once the population analog to Corollary \ref{sample_mean_reg:cor}. \begin{Cor}\label{pop_mean_reg:cor} Suppose that a random pre-shape $Y$ is distributed on $M$ absolutely continuous w.r.t. Riemmanian measure and supported by a tubular neighborhood that admits a slice through $p \in M$. Let $X=\pi\circ Y$ be the corresponding random shape. Then $p$ is regular if $[p]$ is an intrinsic mean of $X$, or if $[p]$ is a Ziezold mean, optimal positioning is invariant and (\ref{Ziez-inj:cond}) is valid. \end{Cor} \section{Extrinsic Means for Kendall's Shape Spaces}\label{ext-means:scn} Let us recall the well known \emph{Veronese-Whitney} embedding for Kendall's planar shape spaces $\Sigma_2^k$. Identify $F_2^k$ with $\mathbb C^{k-1}\setminus \{0\}$ such that every landmark column corresponds to a complex number. This means in particular that $z\in \mathbb C^{k-1}$ is a complex row(!)-vector. With the Hermitian conjugate $a^* = (\overline{a_{kj}})$ of a complex matrix $a=(a_{jk})$ the pre-shape sphere $S_2^k$ is identified with $\{z\in \mathbb C^{k-1}: zz^*=1\}$ on which $SO(2)$ identified with $S^1=\{\lambda \in\mathbb C: |\lambda|=1\}$ acts by complex scalar multiplication. Then the well known Hopf-Fibration mapping to complex projective space gives $\Sigma_2^k=S_2^k/S^1=\mathbb CP^{k-2}$. Moreover, denoting with $M(k-1,k-1,\mathbb C)$ all complex $(k-1)\times (k-1)$ matrices, the Veronese-Whitney embedding is given by \begin{eqnarray*} S_2^k/S^1 &\to& \{a \in M(k-1,k-1,\mathbb C): a^*=a\}\\ ~[z] &\mapsto& z^*z\,. \end{eqnarray*} \begin{Rm}\label{VW-emb-iso:rm} The Procrusets metric of $\Sigma_2^k$ is isometric with the Euclidean metric of $M(k-1,k-1,\mathbb C)$ since we have $\langle z,w \rangle = \operatorname {Re}(zw^*)$ for $z,w\in S_2^k$ and hence, $d^{(p)}([z],[w]) = \sqrt{1-wz^*zw^*} = \|w^*w-z^*z\|/\sqrt{2}$. \end{Rm} The idea of the Veronse-Whitney embedding can be transferred to the general case of shapes of arbitrary dimension $m\geq 2$. Even though the embedding given below is apt only for reflection shape space it can be applied to practical situations in similarity shape analysis whenever the geometrical objects considered have a common orientation. As above, the number $k$ of landmarks is essential and will be considered fixed throughout this section; the dimension $1\leq m < k$, however, is lost in the embedding and needs to be retrieved by projection. To this end recall the (non-unique) embedding of $S_j^k$ in $S_m^k$ $(1\leq j \leq m)$ in (\ref{pre-shape-emb:eq}) which gives rise to a unique and canonical embedding of $R\Sigma_j^m$ in $R\Sigma_m^k$. Moreover, consider the strata $$(R\Sigma_m^k)^j := \{[x]\in R\Sigma_m^k: \operatorname {rank}(x)=j\},~~(\Sigma_m^k)^{j}:=\{[x] \in \Sigma_m^k: \operatorname {rank}(x)=j\}$$ for $j=1,\ldots,m$, each of which carries a canonical manifold structure; due to the above embedding, $(R\Sigma_m^k)^j$ can and will be identified with $(R\Sigma_j^k)^j$ such that $$ R\Sigma_m^k ~=~ \bigcup_{j=1}^m (R\Sigma_j^k)^j\,.$$ At this point we note that $SO(m)$ is connected, while $O(m)$ is not; and the consequences for the respective manifold parts, i.e. points of maximal orbit type: \begin{eqnarray*} (\Sigma_m^k)^* &=& (\Sigma_m^k)^{m-1}\cup (\Sigma_m^k)^m\,,\\%\mbox{ where } (\Sigma_m^k)^{j}:=\{[x] \in \Sigma_m^k: \operatorname {rank}(x)=j\}\,,\\ (R\Sigma_m^k)^* &=& (R\Sigma_m^k)^{m}\,. \end{eqnarray*} Similarly, we have a stratifiction $${\cal P} := \left\{a\in M(k-1,k-1): a=a^T\geq 0, \operatorname {tr}(a)=1\right\} ~=~ \bigcup_{j=1}^{k-1}{\cal P}^j$$ of a bounded flat convex manifold ${\cal P}$ with non-flat unbounded manifolds $${\cal P}^j :=\{a\in {\cal P}: \operatorname {rank}(a) = j\}~ (j=1,\ldots,k-1)\,,$$ all embedded in $M(k-1,k-1)$. The \emph{Schoenberg map} $\mathfrak{s} : R\Sigma_m^k \to {\cal P}$ is then defined on each stratum by $$ \begin{array}{rcl}\mathfrak{s}|_{(R\Sigma_m^k)^j}=:\mathfrak{s}^j: (R\Sigma_m^k)^j &\to& {\cal P}^j \\~[x]&\mapsto&x^Tx\end{array}\,.$$ For $x\in S_j^k$ recall the tangent space decomposition $T_xS_j^k = T_{x}[x]\oplus H_xS_j^k$ into the \emph{vertical tangent space} along the fiber $[x]$ and its orthogonal complement the \emph{horizontal tangent space}. For $x\in (S_j^k)^j$ identify canonically (cf. \citet[p. 109]{KBCL99}): $$ T_{[x]} (R\Sigma_j^k)^j\cong H_xS_j^k = \{w \in M(j,k-1): \operatorname {tr}(wx^T) = 0, wx^T=xw^T\}\,.$$ Then the assertion of the following Theorem condenses results of \cite{BandPat05}, cf. also \cite{DKLW08}. \begin{Th} Each $\mathfrak{s}^j$ is a diffeomorphism with inverse $(\mathfrak{s}^j)^{-1}(a) = [(\sqrt{\lambda} u^T)_1^j]$ where $a = u\lambda u^T$ with $u \in O(k-1)$, $\lambda=\operatorname {diag}(\lambda_1,\ldots,\lambda_m)$, and $0=\lambda_{j+1} =\ldots=\lambda_{k-1}$ in case of $j<k-1$. Here, $(a)_1^j$ denotes the matrix obtained from taking only the first $j$ rows from $a$. For $x\in S_j^k$ and $w\in H_xS_j^k\cong T_{[x]}(R\Sigma_j^k)^j$ the derivative is given by $$d(\mathfrak{s}^j)_{[x]}w = x^Tw+w^Tx\,.$$ \end{Th} \begin{Rm}\label{Schoenberg-emb-non-iso:rm} In contrast to the Vernose-Whitney embedding, the Schoenberg embedding is not isometric as the example of $$ x = \left(\begin{array}{cc}\cos \phi&0\\0&\sin\phi\end{array}\right)\,,~~ w_1 = \left(\begin{array}{cc}\sin \phi&0\\0&-\cos\phi\end{array}\right),~~ w_2 = \left(\begin{array}{cc}0&\cos \phi\\\sin\phi&0\end{array}\right),~~$$ teaches: $$\|x^Tw_1+w_1^Tx\| = \sqrt{2}\, 2|\cos\phi\sin\phi|,~~ \|x^Tw_2+w_2^Tx\| =\sqrt{2}\,.$$ \end{Rm} Since ${\cal P}$ is bounded, convex and Euclidean, the classical expectation $\mathbb E(X^TX)\in {\cal P}^j$ for some $1\leq j\leq k-1$ of the Schoenberg image $X^TX$ of an arbitrary random reflection shape $[X]\in R\Sigma_m^k$ is well defined. Then we have at once the following relation between rank of Euclidean mean and increasing sample size. \begin{Th}\label{dim-Schoenberg-mean:thm} Suppose that a random reflection shape $[X]\in R\Sigma_m^k$ is distribed absolutely continous w.r.t. the projection of the spherical measure on $S_m^k$. Then $$\mathbb E(X^TX)\in {\cal P}^{k-1}\mbox{ and } \frac{1}{n}\sum_{i=1}^n X_i^TX_i\in {\cal P}^{j}~ a.s.$$ for every i.i.d. sample $X_1,\ldots,X_n\sim X$ if $j \leq nm < j+1$ for $nm < k-1$ and $j=k-1$ for $nm \geq k-1$. \end{Th} \begin{figure} \begin{minipage}{0.6\textwidth} \includegraphics[angle=0,width=1\textwidth]{project_lambda_plane1.eps} \end{minipage} \begin{minipage}{0.35\textwidth} \caption{\it Projections (if existent) of two points (crosses) in the $\lambda$-plane to the open line segment $\Lambda=\{(\lambda_1,\lambda_2): \lambda_1+\lambda_2=1, \lambda_1,\lambda_2>0\}$. The dotted line gives the central projections (denoted by stars) suggested by \cite{DKLW08} which is well defined for all symmetric, positive } \end{minipage} {\it definite matrices (corresponding to the first open quadrant), the dashed line gives the orthogonal projection (circle) which is well defined in the triangle below $\Lambda$ (corresponding to ${\cal P}$) and above $\Lambda$ in an open strip. In particular it exists not for the right point.\label{Schoenberg_proj_fig}} \end{figure} Hence, in stastical settings involving a higher number of landmarks, a sufficiently well behaved projection of a high rank Euclidean mean onto lower rank ${\cal P}^r\cong (\Sigma_r^k)^r$, usually $m=r$, is to be employed, giving at once a mean shape satisfying strong consistency and a CLT. Here, unlike to intrinsic or Procrustes analysis, the dimension $r$ chosen is crucial for the dimensionality of the mean obtained. The orthogonal projection $$\begin{array}{rcl}\phi^r : \bigcup_{i=r}^{k-1} {\cal P}^i &\to& {\cal P}^r \\ a&\mapsto&\operatorname {argmin}_{b \in {\cal P}^r} \operatorname {tr}\big((a - b)^2\big)\end{array}$$ giving the set of \emph{extrinsic Schoenberg means} has been computed by \cite{B08}: \begin{Th}\label{og-proj-sb:thm} For $1\leq r\leq k-1$, $a = u\lambda u^T \in {\cal P}$ with $u \in O(k-1)$, $\lambda=\operatorname {diag}(\lambda_1,\ldots,\lambda_m)$, $\lambda_{1}\geq \ldots\geq \lambda_{k-1}$ and $\lambda_r>0$ the orthogonal projection onto ${\cal P}^r$ is given by $$ \phi^r(a) = u \mu u^T$$ with $\mu = \operatorname {diag}(\mu_1,\ldots,\mu_r,0,\ldots,0)$, $$\mu_i = \lambda_i +\frac{1}{r} - \overline{\lambda}_r\,(i=1,\ldots,r)$$ and $\overline{\lambda}_r = \frac{1}{r}\sum_{i=1}^r\lambda_i \leq \frac{1}{r}$ which is uniquely determined if and only if $\lambda_{r}>\lambda_{r+1}$. \end{Th} With the notation of Theorem \ref{og-proj-sb:thm}, a non-orthogonal \emph{central projection} $\psi^r(a) = u \nu u^T$ equally well and uniquely determined has been proposed by \cite{DKLW08} with $\nu = \operatorname {diag}(\nu_1,\ldots,\nu_r,0,\ldots,0)$, $$\nu_i = \frac{\lambda_i}{r\overline{\lambda}_r}~~(i=1,\ldots,r)\,.$$ Orthogonal and central projection are depicted in Figure \ref{Schoenberg_proj_fig}. \section{Local Effects of Curvature}\label{local:scn} In this section we assume that a manifold stratum $M$ supporting a random element $X$ is isometrically embedded in a Euclidean space $\mathbb R^D$ of dimension $D>0$. With the orthogonal projection $\Phi :\mathbb R^D \to M$ from Section \ref{Frechet-rho-means:scn} and the Riemann exponential $\exp_p$ of $M$ at $p$ we have the $$\begin{array}{lcl} \mbox{\emph{intrinsic tangent space coordinate }}&& \exp_p^{-1}X\mbox{ and the}\\ \mbox{\emph{residual tangent space coordinate }}&& d\Phi_p(X-p)\,, \end{array}$$ respectively, of $X$ at $p$, if existent. \subsection{Finite Power of Tests and Tangent Space Coordinates}\label{tangent:scn} With the above setup, assume that $\mu\in M$ is a unique mean of $X$. If $c\in \mathbb R^D$ is the center of the osculatory circle touching the geodesic segment in $M$ from $X$ to $\mu$ at $\mu$ with radius $r$, and if $X_r$ is the orthogonal projection of $X$ to that circle, then $X=X_r + O(\|X-\mu\|^3)$. Moreover, with \begin{eqnarray*} \cos \alpha &=& \left\langle \frac{X-c}{\|X-c\|}, \frac{\mu-c}{r}\right\rangle\\ &=& \frac{1}{r^2}\,\langle X_r-c,\mu-c\rangle + O(\|X-\mu\|^3) \end{eqnarray*} we have the residual tangent space coordinate $$ v = X-c- \frac{\mu-c}{r}\,\|X-c\|\,\cos \alpha = X_r-c -(\mu-c)\cos \alpha+ O(\|X-\mu\|^3)$$ having squared length $\|v\|^2 =r^2\sin\alpha^2 + O(\|X-\mu\|^3)$. By isometry of the embedding, the intrinsic tangent space coordinate is given by $$\exp_{\mu} X = \frac{r\alpha}{\|v\|}\, v + O(\|X-\mu\|^3)\,.$$ With the component $$ \nu = \mu-c - \|X-c\| \frac{\mu-c}{r}\,\cos \alpha = (\mu -c) (1-\cos \alpha) + O(\|X-\mu\|^3) $$ of $X$ normal to the mentioned geodesic segment of squared length $\|\nu\|^2 = r^2(1-\cos\alpha)^2+O(\|X-\mu\|^3)$ we obtain \begin{eqnarray* \|\exp_{\mu} X\|^2 &=& \|v\|^2 + \|\nu\|^2+ O(\|X-\mu\|^3)\,, \end{eqnarray*} since \begin{eqnarray*} (1-\cos\alpha)^2 + \sin^2\alpha &=& 2(1-\cos\alpha) ~=~\alpha^2 + 2\frac{\alpha^4}{4!} + \cdots \end{eqnarray*} and $\alpha = O(\|X-\mu\|)$. In consequence we have \begin{Rm}\label{use_is_ext_not_intr_coord_rm} In approximation, the variation of intrinsic tangent space coordinates is the sum of the variation $\|v\|^2$ of residual tangent space coordinates and the variation normal to it. In particular, for spheres $$ \|\exp_{\mu} X\|^2 ~\geq~ \|v\|^2 + \|\nu\|^2\,.$$ Since the variation in normal space is irrelevant for a two-sample test for equality of means, say, a higher power for tests based on intrinsic means can be expected when solely residual tangent space coordinates obtained from an isometric embedding are used rather than intrinsic tangent space coordinates. Note that the natural tangent space coordinates for Ziezold means are residual. \end{Rm} A simulated classification example in Section \ref{class_data_simulations_scn} illustrates this slight effect. \subsection{The $1:3$ - Property for Spherical and Kendall Shape Means}\label{1:3:scn} In this section $M=S^{D-1}\subset \mathbb R^D$ is the $(D-1)$-dimensional unit-hypersphere embedded isometrically in Euclidean $D$-dimensional space. The orthogonal projection $\Phi: \mathbb R^D \to S^{D-1}: p \to \frac{p}{\|p\|}$ is well defined except for the origin $p=0$, and the normal space at $p\in S^{D-1}$ is spanned by $p$ itself. In consequence, a random point $X$ on $S^{D-1}$ has $$d\Phi_p(X-p) = X -p \cos \alpha,~~\exp_p(X) = \frac{\alpha}{\sin \alpha}~ d\Phi_p(X-p)$$ as residual and intrinsic, resp., tangent space coordinate at $-X\neq p\in S^{D-1}$ where $\cos\alpha = \langle X,p\rangle$. \begin{Rm} An adaption (i.e. multiplying by the respective $\frac{\sin \alpha}{\alpha}$ in every iteration step) of the algorithm for intrinsic means in \citet[Section 5.4]{HZ06} gives at once an algorithm converging usually quickly to one of the two sample residual means. \end{Rm} \begin{Th}\label{unique-intr-mean:th} If $X$ is contained in a closed half sphere with non-zero probability in its interior, it has a unique intrinsic mean which is assumed in the interior of that half sphere. \end{Th} \begin{proof} Below, we show that every intrinsic mean necessarily lies within the interior of the half sphere. Then, \citet[Theorem 7.3]{KWS90} yields uniqueness. W.l.o.g. let $X = (\sin\phi, x_2,\ldots, x_n)$ such that ${\cal P}\{\sin\phi <0\} = 0 < {\cal P}\{\sin\phi >0\}$ and assume that $p = (\sin \psi, p_2,\ldots, p_n)\in S^{D-1}$ is an intrinsic mean. Moreover let $p' = (\sin (|\psi|), p_2,\ldots, p_n)$. Since \begin{eqnarray*} \operatorname {\mathbb E}\left(\|\exp_p(X)\|^2\right) &=& \operatorname {\mathbb E}\left(\arccos^2 \langle p,X\rangle\right)\\ &=&\operatorname {\mathbb E}\left(\arccos^2 \left(\sin\psi\sin\phi + \sum_{j=2}^np_jx_j\right)\right)\\ &\leq& \operatorname {\mathbb E}\left(\|\exp_{p'}(X)\|^2\right) \end{eqnarray*} with equality if and only if $|\psi| =\psi$, this can only happen for $\psi\geq 0$. Now, suppose that $p=(0,p_2,\ldots,p_n)$ is an intrinsic mean. For deterministic $\psi \geq 0 $ consider $p(\psi) = (\sin \psi,p_1\cos\psi,\ldots,p_n\cos\psi)$. Then \begin{eqnarray*} \operatorname {\mathbb E}\left(\|\exp_{p(\psi)}(X)\|^2\right) &=&\operatorname {\mathbb E}\left(\arccos^2 \left(\sin\psi\sin\phi + \cos \phi \sum_{j=2}^np_jx_j\right)\right)\\ &=& \operatorname {\mathbb E}\left(\|\exp_{p}(X)\|^2\right) - C_1 \psi + O(\psi^2) \end{eqnarray*} with $C_1 >0$ since ${\cal P}\{\sin\phi >0\}>0$. In consequence, $p$ cannot be an intrinsic mean. Hence, we have shown that every intrinsic mean is contained in the interior of the half sphere. \end{proof} \begin{Rm}\label{Karcher-means:rm} For the special case of spheres, this extends a result of H. Le, asserting a unique intrinsic mean on general Riemannian manifolds, if among others, the support of $X$ is contained in a geodesic quarter-ball, see \citet{L01,L04}. H. Le's result is an extension of \citet{Ka77}, asserting uniqueness of intrinsic means in geodesic balls if among others, $X$ is supported in a geodesic quarter ball. Intrinsic means restricted to geodesic balls are often called \emph{Karcher means}. As mentioned in the above proof, \citet[Theorem 7.3]{KWS90} extended Karcher's result to obtain uniquenes of Karcher means, if among others, the support of $X$ is contained in a half ball. In view of Theorem \ref{unique-intr-mean:th}, one might expect to extend Le's result on uniquenes of intrinsic means on the entire manifold accordingly, to random elements contained in a geodesic half ball. \end{Rm} The following Theorem characterizes the three spherical means. \begin{Th}\label{conditions_for_spherical_means_th} Let $X$ be a random point on $S^{D-1}$. Then $x^{(e)} \in S^{D-1}$ is the unique extrinsic mean if and only if the Euclidean mean $\mathbb E(X) = \int_{S^{D-1}} X\,dP_X$ is non-zero. If the following right hand sides are non-zero, then there are suitable $\lambda^{(e)} = \|\mathbb E(X)\|>0$, $\lambda^{(r)}>0$ and $\lambda^{(i)}>0$ such that $$\lambda^{(e)} x^{(e)} = \mathbb E(X)\,,$$ every residual mean $x^{(r)} \in S^{D-1}$ satisfies $$\lambda^{(r)} x^{(r)} = \mathbb E\big(\langle X,x^{(r)}\rangle\, X\big)\,,$$ and every intrinsic mean $x^{(i)} \in S^{D-1}$ satisfies $$\lambda^{(i)} x^{(i)} = \mathbb E\left(\frac{\arccos\langle X,x^{(i)}\rangle}{\sqrt{1-\langle X,x^{(i)}\rangle^2 }}\, X\right)\,.$$ In the last case we additionally require that $\operatorname {\mathbb E}\left(\frac{\arccos\langle X,x^{(i)}\rangle}{\sqrt{1-\langle X,x^{(i)}\rangle^2 }}\, \langle X,x^{(i)}\rangle\right)> 0$ which is in particular the case if $X$ is contained in a closed half sphere with non-zero probability in the interior. \end{Th} \begin{proof} The assertions for the extrinsic mean are well known from \cite{HLR96}. The second assertion for residual means follows from minimization of $$ \int_{S^{D-1}} \|p-\langle p,x\rangle x\|^2\,dP_X(p) = 1 - \int_{S^{D-1}} \langle p,x\rangle^2\,dP_X(p)$$ with respect to $x\in \mathbb R^D$ under the constraining condition $\|x\|=1$. Using a Lagrange ansatz this leads to the necessary condition $$ \int_{S^{D-1}} \langle p,x\rangle\, p\,dP_X(p) = \lambda x$$ with a Lagrange multiplier $\lambda$ of value $\operatorname {\mathbb E}(\langle X,x\rangle^2)$ which is positive unless $X$ is supported by the hypersphere orthogonal to $x$. In that case, by Theorem \ref{res_means_sphere:rm}, $x$ cannot be a residual mean of $X$, as every residual mean is as well contained in that hypersphere. Hence, we have $\lambda >0$. The same method applied to $$ \int_{S^{D-1}} \|\exp^{-1}_x(p)\|^2 \,dP_X(p) = \int_{S^{D-1}} \arccos^2(\langle p,x\rangle)\,dP_X(p) $$ taking into account Theorem \ref{unique-intr-mean:th} yields the third assertion on the intrisic mean. \end{proof} In particular, residual means are eigenvectors to the largest eigenvalue of the matrix $\mathbb E(XX^T)$. As such, they rather reflect the mode then the classical mean of a distribution: \begin{Ex}\label{different_means_ex} Consider $\gamma \in (0,\pi)$ and a random variable $X$ on the unit circle $\{e^{i\theta}: \theta \in [0,2\pi)\}$ which takes the value $1$ with probability $2/3$ and $e^{i\gamma}$ with probability $1/3$. Then, explicit computation gives the unique intrinsic and extrinsic mean as well as the two residual means $$ x^{(i)} = e^{i\frac{\gamma}{3}},\quad x^{(e)} = e^{i\arctan\frac{\sin \gamma}{2+\cos\gamma}},\quad x^{(r)} = \pm\, e^{i\frac{1}{2}\,\arctan\frac{\sin (2\gamma)}{2+\cos(2\gamma)}}\,.$$ Figure \ref{different_means_fig} shows the case $\gamma = \frac{\pi}{2}$. \end{Ex} \begin{figure} \begin{minipage}{0.5\textwidth} \includegraphics[width=0.95\textwidth]{different_means_S1.eps} \end{minipage} \begin{minipage}{0.5\textwidth} \caption{\it Means on a circle of a distribution taking the upper dotted value with probability $1/3$ and the lower right dotted value with probability $2/3$. The latter happens to be one of the two residual means.\label{different_means_fig}} \end{minipage} \end{figure} In contrast to Figure \ref{different_means_fig}, one may assume in many practical applications that the mutual proximities of the unique intrinsic mean $x^{(i)}$, the unique extrinsic mean $x^{(e)}$ and the unique residual mean $x^{(r_0)}$ closer to $x^{(e)}$ are rather small, namely of the same order as the squared proximity of the modulus $\|\mathbb E(X)\|$ of the Euclidean mean to $1$, cf. Table \ref{1:3:tab}. We will use the following condition \begin{eqnarray}\label{reg_extr_mean_cond}\left.\begin{array}{rcl} \|x^{(e)} -x^{(r_0)}\|\,,~~ \|x^{(e)} -x^{(i)}\|&=& O\big((1-\|\operatorname {\mathbb E}(X)\|)^2\big) \end{array}\right. \end{eqnarray} with the \emph{concentration parameter} $1-\|\operatorname {\mathbb E}(X)\|$. \begin{Cor}\label{3_means_on_circle_cor} Under condition (\ref{reg_extr_mean_cond}), if all three means are unique, then the great circular segment between the residual $x^{(r_0)}$ mean closer to the extrinsic mean $x^{(e)}$ and the intrinsic mean $x^{(i)}$ is divided by the extrinsic mean in approximation by the ratio $1:3$: \begin{eqnarray* x^{(r_0)} &=& \frac{\|\operatorname {\mathbb E}(X)\|}{\lambda^{(r_0)}}\left(x^{(e)} - \frac{\mathbb E\big( \langle X-x^{(e)} ,X\rangle\,X\big)}{\|\operatorname {\mathbb E}(X)\|} + O\big((1-\|\operatorname {\mathbb E}(X)\|)^2\big)\right)\\ x^{(i)} &=& \frac{\|\operatorname {\mathbb E}(X)\|}{\lambda^{(i)}}\left(x^{(e)} + \frac{1}{3}\, \frac{\mathbb E\big( \langle X-x^{(e)} ,X\rangle\,X\big)}{\|\operatorname {\mathbb E}(X)\|} + O\big((1-\|\operatorname {\mathbb E}(X)\|)^2\big)\right)\, \end{eqnarray*} with $\lambda^{(i)}$ and $\lambda^{(r_0)}$ from Theorem \ref{conditions_for_spherical_means_th}. \end{Cor} \begin{proof} For any $x,p \in S^{D-1}$ decompose $ p - x = p - \langle x,p\rangle\, x - z(x,p) x $ with $z(x,p) = 1 - \langle x,p\rangle$, the length of the part of $p$ normal to the tangent space at $x$. Note that $\operatorname {\mathbb E}\big(z(x^{(e)},X\big) = 1-\|\operatorname {\mathbb E}(X)\|$. Now, under condition (\ref{reg_extr_mean_cond}), verify the first assertion using Theorem \ref{conditions_for_spherical_means_th}: \begin{eqnarray*}\label{approx_extr_res_mean}\nonumber x^{(r0)} &=&\frac{1}{\lambda^{(r_0)}}\left(\operatorname {\mathbb E}(X) - \operatorname {\mathbb E}\big(z(x^{(r_0)},X)X\big)\right)\,. \end{eqnarray*} On the other hand since \begin{eqnarray*} \frac{\arccos (1-z)}{\sqrt{1-(1-z)^2 }} &=& 1 + \frac{1}{3}\,z +\frac{2}{15}\,z^2 +\ldots \end{eqnarray*} we obtain with the same argument the second assertion \begin{eqnarray*} x^{(i)} &=& \frac{1}{\lambda^{(i)}} \operatorname {\mathbb E} \left(\frac{\arccos\langle X,x^{(i)}\rangle}{\sqrt{1-\langle X,x^{(i)}\rangle^2 }}\, X\right)\\ &=& \frac{1}{\lambda^{(i)}}\left(\operatorname {\mathbb E}(X)+ \frac{1}{3}\, \operatorname {\mathbb E}\left( z(x^{(i)},X)\,X\right) + \frac{2}{15}\, \operatorname {\mathbb E}\left( z(x^{(i)},X)^2\,X\right) +\ldots\right)\\ &=& \frac{\|\operatorname {\mathbb E}(X)\|}{\lambda^{(i)}}\left(x^{(e)} +\frac{1}{3\|\operatorname {\mathbb E}(X)\|}\, \operatorname {\mathbb E} \left(z(x^{(e)},X)\,X\right) + O\big((1-\|\operatorname {\mathbb E}(X)\|)^2\big)\right)\,. \end{eqnarray*} \end{proof} \begin{Rm} The tangent vector defining the great circle approximately connecting the three means is obtained from correcting with the expected normal component of any of the means. As numerical experiments show, this great circle is different from the first \emph{principal component geodesic} as defined in \cite{HZ06}. \end{Rm} Recall the following connection between intrinsic and extrinsic spherical versus intrinsic and Ziezold shape means, respectively, cf. Lemmas \ref{sample_int_mean_bottom_int_mean_top:lem} and \ref{pop_int_mean_bottom_int_mean_top:lem}. It is easy to see the corresponding connection for residual versus Procrustes means. \begin{Rm} Suppose that a random shape $X$ on $\Sigma_m^k$ is supported by $R\subset \Sigma_m^k$ admitting a horizontal measurable lift $L\subset S_m^k$ in optimal position to $p\in S_m^k$. Define the random variable $Y$ on $S_m^k$ by $\pi \circ Y = X$. Then we have that \begin{enumerate} \item[] if $[p]$ is an intrinsic mean of $X$ then $p$ is an intrinsic mean of $Y$, \item[] if $[p]$ is a Procrustes mean of $X$ then $p$ is a residual mean of $Y$, \item[] if $[p]$ is a Ziezold mean of $X$ then $p$ is an extrinsic mean of $Y$ \end{enumerate} \end{Rm} In consequence, Corollary \ref{3_means_on_circle_cor} extends at once to Kendall's shape spaces. \begin{Cor}\label{3_means_on_kendall_cor} Suppose that a random shape $X$ on $\Sigma_m^k$ with unique intrinsic mean $\mu^{(i)}$, unique Ziezold mean $\mu^{(z)}$ and unique Procrustes mean $\mu^{(p_0)}$ closer to $\mu^{(i)}$ is supported by $R\subset \Sigma_m^k$ admitting a horizontal measurable lift $L\subset S_m^k$ in optimal position to $\mu^{(z)} \in S_m^k$. If the means are sufficiently close to each other in the sense of $$d_{\Sigma_m^k}(\mu^{(z)} -\mu^{(p_0)})\,,~~ d_{\Sigma_m^k}(\mu^{(z)} -\mu^{(i)})~=~ O\big((1-\|\operatorname {\mathbb E}(Y)\|)^2\big) $$ with the random pre-shape $Y$ on $L$ defined by $X = \pi \circ Y$, then the generalized geodesic segment between $\mu^{(i)}$ and $\mu^{(p_0)}$ is approximately divided by $\mu^{(z)}$ by the ratio $1:3$ with an error of order $O\big((1-\|\operatorname {\mathbb E}(Y)\|)^2\big)$. \end{Cor} \section{Examples: Exemplary Datasets and Simulations}\label{class_data_simulations_scn} \subsection{The $1:3$ property} In the first example we illustrate Corollary \ref{3_means_on_kendall_cor} on the basis of four data sets: \begin{description} \item[poplar leaves:] contains 104 quadrangular planar shapes extracted from poplar leaves in a joint collaboration with Institute for Forest Biometry and Informatics at the University of G\"ottingen, cf. \citet{HHM09}. \item[digits '3':] contains 30 planar shapes with 13 landmarks each, extracted from handwritten digits '3', cf. \citet[p. 318]{DM98}. \item[macaque skulls:] contains three-dimensional shapes with 7 landmarks each, of 18 macaque skulls, cf. \citet[p. 16]{DM98}. \item[iron age brooches:] contains 28 three-dimensional tetrahedral shapes of iron age brooches, cf. \citet[Section 3.5]{S96}. \end{description} \begin{figure}[h!]\centering \includegraphics[angle=-90,width=0.45\textwidth]{three_means_leaves.eps} \includegraphics[angle=-90,width=0.45\textwidth]{three_means_digit3.eps} \\ \includegraphics[angle=-90,width=0.45\textwidth]{three_means_macq.eps} \includegraphics[angle=-90,width=0.45\textwidth]{three_means_brooch.eps} \caption{\it Depicting shape means for four typical data sets: intrinsic (star), Ziezold (circle) and Procrustes (diamond) projected to the tangent space at the intrinsic mean. The cross divides the generalized geodesic segment joining the intrinsic with the Procrustes mean by the ratio $1:3$.\label{three_means:fig}} \end{figure} \begin{table}[h! { \footnotesize $$\begin{array}{r|ccc|l} \mbox{data set &d_{\Sigma_m^k}(\mu^{(i)},\mu^{(z)})&d_{\Sigma_m^k}(\mu^{(p_0)},\mu^{(z)})&d_{\Sigma_m^k}(\mu^{(p_0)},\mu^{(i)})&(1-|\operatorname {\mathbb E}(Y)|)^2\\ \hline \mbox{poplar leaves}&6.05e-05&1.83e-04&2.44e-04&5.24e-05\\ \mbox{digits '3'}&0.00154&0.00452&0.00605& 0.00155\\ \mbox{macaque skulls}&1.96e-05&5.89e-05&7.85e-05&7.59e-06\\ \mbox{iron age brooches}&0.000578&0.001713&0.002291&0.000217 \end{array}$$} \caption{\it Mutual shape distances between intrinsic mean $\mu^{(i)}$, Ziezold mean $\mu^{(z)}$ and Procrustes mean $\mu^{(p_0)}$ closer to the Ziezold mean for various data sets. Right column: the squared distance between modulus of Euclidean mean and $1$, cf. Corollary \ref{3_means_on_kendall_cor}.\label{1:3:tab}} \end{table} As clearly visible from Figure \ref{three_means:fig} and Table \ref{1:3:tab}, the approximation of Corollary \ref{3_means_on_kendall_cor} for two- and three-dimensional shapes is highly accurate for data of little dispersion (the macaque skull data) and still fairly accurate for highly dispersed data (the digits '3' data). \begin{figure}[h!] \includegraphics[angle=-90,width=1\textwidth]{procrustes_blind.eps} \caption{\it A data set of three planar triangles (top row) with its corresponding intrinsic mean (bottom left), Ziezold mean (bottom center) and Procrustes mean (bottom right).\label{procrustes_blind:fig}} \end{figure} \begin{figure}[h!] \centering \includegraphics[angle=-90,width=0.3\textwidth]{schoen_sampl1.eps} \includegraphics[angle=-90,width=0.3\textwidth]{schoen_sampl2.eps} \includegraphics[angle=-90,width=0.3\textwidth]{schoen_sampl3.eps}\\ \includegraphics[angle=-90,width=1\textwidth]{schoen_blind.eps} \caption{\it Planar triangles $q_1= x\cos\beta - w_2\sin\beta$, $q_2= x\cos\beta - w_2\sin\beta$ and $q = x\cos\beta' + w_1\sin\beta'$ with $x,w_1,w_2$ from Remark \ref{Schoenberg-emb-non-iso:rm}, $\phi =0.05$, $\beta=0.3=\beta'$ (top row). Intrinsic means (middle row) of sample $(q_1,q_2)$ (left) and $(q_1,q_2,q)$ (right). Schoenberg means (bottom row) of sample $(q_1,q_2)$ (left) and $(q_1,q_2,q)$ (right).\label{Schoenberg_blind:fig} } \end{figure} \subsection{``Blindness'' of Procrustes and Schoenberg Means} In the second example we illustrate an effect of ``blindness to data'' of Procrustes means and Schoenberg means. The former blindness is due to the affinity of the Procrustes mean to the mode in conjunction with curvature, the latter is due to non-isometry of the Schoenberg embedding. While the former effect occurs only for some highly dispersed data when the analog of condition (\ref{reg_extr_mean_cond}) is violated, the latter effect is local in nature and may occur for concentrated data as well. \begin{Ex} Reenacting the situation of Example \ref{different_means_ex} and Figure \ref{different_means_fig}, the shapes of the triangles $P$ and $Q$ in Figure \ref{procrustes_blind:fig} are almost maximally remote. Since the mode $P$ is assumed twice and $Q$ only once, the Procrustes mean is nearly blind to $Q$. In case of $$ P=\left(\begin{array}{cccc}1&-1&0&0\\0&0&0&0\\0&0&0&0\end{array}\right),~~ Q= \left(\begin{array}{cccc}1&1&-2&0\\\frac{1}{\sqrt{2}}&\frac{1}{\sqrt{2}}&\frac{1}{\sqrt{2}}&-\,\frac{3}{\sqrt{2}}\\0&0&0&0\end{array}\right)$$ the Procrustes mean shape of the sample $P,P,Q$ is the one-dimensional $P$ because of blindness to the two-dimensional $Q$. \end{Ex} Let us record this fact. \begin{Rm}\label{Procrustes_deg:rm} The Procrustes mean of a random shape assuming both regular and degenerate shapes with non-zero probability may be degenerate. \end{Rm} Even though Schoenberg means have been introduced to tackle 3D shapes, the effect of ``blindness'' can be well illustrated already for 2D. To this end consider $x=x(\phi)$, $w_1=w_1(\phi)$ and $w_2=w_2(\phi)$ as introduced in Remark \ref{Schoenberg-emb-non-iso:rm}. Along the horizontal geodesic through $x$ with initial velocity $w_2$ we pick two points $q_1= x\cos\beta +w_2\sin\beta$ and $q_2= x\cos\beta - w_2\sin\beta$. On the orthogonal horizontal geodesic through $x$ with initial velocity $w_1$ pick $q = x\cos\beta' + w_1\sin\beta'$. Recall from Remark \ref{Schoenberg-emb-non-iso:rm}, that along that geodesic the derivative of the Schoenberg embedding can be made arbitrarily small for $\phi$ near $0$. Indeed, Figure \ref{Schoenberg_blind:fig} illustrates that for small $\phi$, the Schoenberg mean is nearly unchanged if the triangle $q$ is added to the sample $q_1,q_2$. The corresponding intrinsic means, however, are sensitive to the sample's expansions. \begin{figure}[h!] \centering \includegraphics[width=0.45\textwidth]{cube.eps} \includegraphics[width=0.45\textwidth]{pyramid.eps} \caption{\it cube (left) and pyramid (right) for classification.\label{cube_pyramid:fig}} \end{figure} \subsection{Classification Power of a Test} In the ultimate example we illustrate the consequences of the choice of tangent space coordinates and the effect of the tendency of the Schoenberg mean to increase dimension by a classification simulation. To this end we apply a Hotelling $T^2$-test to discriminate the shapes of 10 noisy samples of regular unit cubes from the shapes of 10 noisy samples of pyramids with top section chopped off, each with 8 landmarks, given by the following configuration matrix $$\left(\begin{array}{cccccccc} 0& 1 &\frac{1+\epsilon}{2}& \frac{1-\epsilon}{2}& 0& 1&\frac{1+\epsilon}{2}& \frac{1-\epsilon}{2}\\ 0& 0& \frac{1-\epsilon}{2}&\frac{1-\epsilon}{2} & 1& 1& \frac{1+\epsilon}{2}&\frac{1+\epsilon}{2}\\ 0& 0& \epsilon& \epsilon& 0& 0&\epsilon&\epsilon\end{array} \right)\,$$ (cf. Figure \ref{cube_pyramid:fig}) determined by $\epsilon >0$. In the simulation, indenpendent Gaussian noise with variance $\sigma^2>0$ is added to each landmark measurement. Table \ref{sim-tab} gives the number of correct classifications for 1,000 simulations for given $\alpha$, $\epsilon$ and significance level. \begin{table}[h!]\centering\fbox{\footnotesize $\begin{array}{rrr|ccccc} \sigma&\epsilon&\mbox{level}&\multicolumn{2}{c}{\mbox{intrinsic mean with}}&\mbox{Ziezold}&\mbox{Procrustes}&\mbox{Schoenberg}\\ &&&\mbox{residual}&\mbox{intrinsic}&\mbox{mean}&\mbox{mean}&\mbox{mean}\\ &&&\multicolumn{2}{c}{\mbox{tangent space coordinates}}\\ \hline 0.35 & 0.05 & 0.1 & 522 & 542 & 538 & 531 & 442\\ 0.3 & 0.1 & 0.1 &558 & 583 & 581 & 575 & 483\\ 0.2 & 0.2 & 0.05 &556 & 575 & 574 & 575 & 512\\ 0.1 & 0.3 & 0.01 &359 & 369 & 369 & 368 & 366 \\ 0.05 & 0.35 & 0.01 &764 & 780 & 780 & 780 & 809 \end{array}$} \caption{\it Number of correct classifications within 1,000 simulations each of $10$ unit-cubes and pyramids determined by $\epsilon$ (which gives the height), where each landmark is independently corrupted by Gaussian noise of variance $\sigma^2$, via a Hotelling T$^2$-test for equality of means to the given significance level.\label{sim-tab}} \end{table} As visible from Table \ref{sim-tab}, discriminating nearly flat pyramids ($\epsilon = 0.05$) from cubes is achieved much better by employing intrinsic, Ziezold or Procrustes means rather than Schoenberg means. This finding is in concord with Theorem \ref{dim-Schoenberg-mean:thm}: samples of size $10$ of three-dimensional configurations yield Euclidean means a.s. in ${\cal P}^{7}$ which are projected to ${\cal P}^{3}$ to obtain Schoenberg means in $\Sigma_3^8$. In consequence, Schoenberg means of nearly two-dimensional pyramids are essentially three dimensional. With increased height of the pyramid ($\epsilon=0.35$), i.e. for more pronounced third dimension and increased proximity to the unit cube), all means perform almost equally well, with a tendency of the Schoenberg mean to eventually outperform the others. Moreover in any case, intrinsic means with intrinsic tangent space coordinates behave much poorer in view of shape discrimination than intrinsic means with residual tangent space coordinates, cf. Remark \ref{use_is_ext_not_intr_coord_rm}. The latter (intrinsic means with residual tangent space coordinates) are better or equally good as Ziezold and Procrustes means (which naturally use residual tangent space coordinates). \begin{table}[h!]\centering \fbox{ \begin{tabular}{cccc} intrinsic mean&Ziezold mean& Procrustes mean &Schoenberg mean\\\hline $235.167$&$ 182.335 $&$ 225.982 $&$ 35.595$\end{tabular}} \caption{\it Average time in seconds for computation of $1,000$ means of sample size $20$.\label{comp_time:tab}} \end{table} In conclusion we record the time for mean computation in Table \ref{comp_time:tab}. While Ziezold means compute in approximately $3/4$ of the computational time for intrinsic means, Schoenberg means are obtained approximately $5$ times faster. Note that we have included inference based on Procrustes means for illustration, even though no stability result is available and hence at this point, it cannot be theoretically justified that the CLT holds for Procrustes means as well. \section{Discussion}\label{discussion-sec} By establishing stability results for intrinsic and Ziezold means on the manifold part of a shape space, a gap in asymptotic theory for general non-manifold shape spaces could be closed, now allowing for multi-sample tests of equality of intrinsic means and Ziezold means. A similar stability assertion in general is false for Procrustes means. Note that the argument applied to intrinsic and Ziezold means fails for Procrustes means, since in contrast to (\ref{Ziez-inj:cond}) the sum of Procrustes residuals is in general non-zero. The findings on dimensionality condense to \begin{center} Procrustes means may decrease dimension,\\ intrinsic and Ziezold means preserve dimension,\\ Schoenberg means tend to increase dimension. \end{center} Due to the proximity of Ziezold and intrinsic means on Kendall's shape spaces in most practical applications, taking into account that the former are computationally easier accessable (optimally positioning and Euclidean averaging in every iteration step) than intrinsic means (optimally positioning and weighted averaging in every iteration step), in practical applications Ziezold means, may be preferred over intrinsic means. They may be even more preferred over intrinsic means, since Ziezold means naturally come with residual tangent space coordinates which may allow in case of intrinsic means for a slightly higher finite power of tests than intrinsic tangent space coordinates. Computationally much faster (not relying on iteration at all) are Schoenberg means which are available for Kendall's reflection shape spaces. As a drawback, however, Schoneberg means, seem less sensitive for dimensionality of configurations considered than intrinsic or Ziezold means. In consequence, multi-sample test based on Schoenberg means may have a considerably lower power than when based on intrinsic or Ziezold means. We note that Ziezold means may be defined for the shape spaces of planar curves introduced by \cite{ZR72}, which are currently very popular e.g. \cite{KSMJ04} or \cite{SCC06}. Employing Ziezold means there, a computational advantage greater than found here can be expected since computation of iterates of intrinsic means involve computations of geodesics which themselves can only be found iteratively. \section*{Acknowledgment} The author would like to thank Alexander Lytchak for helpful advice on differential geometic issues. Also, the author gratefully acknowledges support by DFG grant MU 1230/10-1 and GRK 1023. \bibliographystyle{../../BIB/elsart-harv}
\section{Introduction} \label{sec:introduction} At PLACES'08, we discussed the need to investigate benchmark examples of session types \cite{DBLP:conf/parle/TakeuchiHK94,DBLP:conf/esop/HondaVK98} to compare productivity, safety and performance with other communications programming languages. As a starting point into the investigation of these issues, we examine SJ \cite{SJwebsite}, the first full object-oriented language to incorporate session types for type-safe concurrent and distributed programming. The SJ language extends Java with syntax for declaring session types (protocols), and a set of core operations (session initiation, send/receive) and high-level constructs (branching, iteration, recursion) for implementing the interactions that comprise the sessions. The SJ compiler statically verifies session implementations against their declared types. Together with runtime compatibility validation between peers at session initiation, SJ guarantees communication safety in terms of message types and the structure of interaction. SJ has been shown to perform competitively with widely-used communication APIs such as network sockets, in certain cases out-performing RMI \cite{SJecoop08}. This paper reports our on-going work on implementing parallel algorithms in SJ, with focus on the aforementioned aspects: {\bf\em productivity} (including code readability and writability), {\bf\em safety} (freedom from type and communication errors \cite{DBLP:conf/parle/TakeuchiHK94,DBLP:conf/esop/HondaVK98}), and {\bf\em performance} (optimisations enabled by SJ, and comparison against other communication systems). Parallel algorithms is a prominent topic in algorithmic research due to the increase of hardware resources such as multicore machines and clusters. The session-based programming methodology and expressiveness of SJ are demonstrated through implementations of: (1) a Monte Carlo approximation of $\pi$, (2) the Jacobi solution of the Discrete Poisson Equation, and (3) a simulation of the $n$-Body problem. These algorithms were selected to evaluate the SJ representation of, amongst other features, typical {\em task and data decomposition} patterns \cite{patternsforparallelalgo} (as featured in 1 and 2), a technique for exchanging {\em ghost points} \cite{usingmpi} (in 2), and an intricate communication pattern over a circular pipeline structure (3). SJ is an evolving framework, and recent extensions to the SJ language \cite{SJecoop08} (e.g. new multicast output operations and advanced iteration structures) and the SJ Runtime (e.g.~improved extensibility through the Abstract Transport) play an important part in the implementation of these algorithms. Using these programs, which feature complex and representative interaction structures, we contribute new benchmark results for analysis to supplement the existing benchmarks for SJ. In particular, benchmark comparisons between SJ and MPJ Express \cite{mpjexpress}, a reference Java messaging system based on the MPI \cite{MPI} standard, for (1) and (2) yield further promising performance results for SJ. We also show how SJ {\em noalias} types can greatly optimise performance, such as for the shared memory communication of the ghost points in (2). We then compare the SJ implementations of the above algorithms with their MPI counterparts from programming perspectives. Despite rich libraries and functionality, MPI remains a low-level API, and can suffer from such commonly perceived disadvantages of explicit message passing as unexpected message structures and deadlocks due to incorrect protocol implementations. From our experiences implementing the above algorithms, we found high-level session programming to be easier than the basic MPI functions, which often require manipulating numerical process identifiers and array indexes (e.g.~for message lengths in (3)) in tricky ways. SJ is able to exploit session types to compensate for, or eliminate, many of the MPI problems: session types themselves are inherently deadlock free, for example. In conclusion, we observe that high-level session abstraction has significant impact on program structure, improving readability and reliability, and session type-safety can greatly facilitate the task of communications programming whilst retaining competitive performance. We also argue that extending SJ with full multiparty session types would allow richer topologies such as the ring and 2D-mesh to be expressed more naturally, and enable performance improvements through massive parallelism. \section{Monte Carlo $\pi$ Approximation} \label{sec:pi} A simple Monte Carlo simulation for approximating the value of $\pi$ is amenable to parallelisation. We use this example to (1) introduce basic and some new SJ constructs; (2) show their use in the description of a simple task decomposition pattern \cite{patternsforparallelalgo}; and (3) demonstrate the effect of parallelisation for performance gain in SJ (\S~\ref{sec:bmarks}). A unit square inscribes a circle of area $\pi/4$; hence, $\pi = 4t$, where $t$ is the ratio of the circle area to the square. $t$ can be determined by selecting a random set of points within the square ($(x, y)$ where $x, y \in [-1, 1]$), and checking how many fall inside the inscribed circle ($x^2 + y^2 <= 1$). A Master process (or thread) can instruct Workers to independently generate and check multiple sets of points in parallel, calculating the final value by combining the results from each Worker. The simple session type, from the Worker side, for the communications involved is: \begin{center} \SJ{protocol workerToMaster \{ sbegin.?(int).!<int> \}} \end{center} Each Worker service (\SJ{sbegin}) is told how many points to test by the Master (\SJ{?(int)}) and sends back the number that fall inside the circle (\SJ{!<int>}). The code for a basic SJ implementation looks like \vspace{1mm} \noindent \begin{tabular}{ll} \begin{lstlisting} // Workers run the simulation. int trials = s_wm.receive(); // ?(int) for(int i = 0; i < trials; i++) if(hit()) hits++; s_wm.send(hits); // !<int> \end{lstlisting} & \begin{lstlisting} // Master controls the Workers. <s_mw1, s_mw2, ...>.send(trials); // Multicast. int totalHits = // Collect the results. s_mw1.receive() + s_mw2.receive() + ...; \end{lstlisting} \end{tabular} \\ \vspace{1mm} \noindent where \SJ{s_mw1} is the Master's session socket to Worker1, etc.; \SJ{s_wm} a Worker's session with the Master; and \SJ{hit} returns the boolean from testing a generated point. The Master can then calculate $t$ by \SJ{totalHits / (trials * n)}, where \SJ{n} is the number of Workers. The SJ compiler statically verifies correctness by checking each session implementation against its declared type (e.g. \SJ{s_wm} against \SJ{workerToMaster}). Then at runtime, session initiation validates the session types of each peer to ensure \emph{duality} between the peers. If successful, the session is established; otherwise, both parties raise an \SJ{SJIncompatibleSessionEception} and the session is aborted. The SJ Runtime is also responsible for failure handling during session execution: if an error occurs at one session peer, e.g. an exception is raised, the failure signal is propagated to all relevant session parties, maintaining consistency across dependent sessions; see \cite{SJecoop08} for more detailed explanation. \section{Jacobi Solution of the Discrete Poisson Equation} \label{sec:jacobi} The implementation of this algorithm demonstrates (1) the expressiveness of SJ due to multicast session-iteration operation; (2) guaranteed type and communication safety in SJ; (3) a type-directed optimisation (for exchanging ghost points) using the new SJ {\em noalias} type; and (4) the {\em transport-independence} of SJ programs, due to the design of the SJ language-runtime framework. Poisson's Equation is a partial differential equation with applications in, for example, heat flow, electrostatics, gravity and climate computations. The discrete two-dimensional Poisson equation $(\nabla^2u)_{ij}$ for a $m \times n$ grid can be written as the formula in (a), \vspace{4mm} \begin{figure}[h] \centering (a)~ $u_{ij} = \frac{1}{4}(u_{i-1, j} + u_{i+1, j} + u_{i, j-1} + u_{i, j+1} - dx^2g_{i, j})$ \hspace{6mm} (b)~ $u_{ij}^{k+1} = \frac{1}{4}(u_{i+1, j}^k + u_{i-1, j}^k + u_{i,j+1}^k + u_{i, j-1}^k)$ \label{fig:jacobi} \vspace{4mm} \end{figure} \noindent where $2 \leq i \leq m - 1$, $2 \leq j \leq n - 1$, and $dx = 1 / (n + 1)$. Jacobi's Method converges on a solution by repeatedly replacing each element of the matrix $u$ by an average of its four neighbouring values and $dx^2g_{i, j}$. For this example, we set $g$ to $0$; then from the $k$-th approximation of $u$, the next iteration performs the calculation in (b) above. \noindent Termination may be on reaching a target convergence threshold or completing a certain number of iterations. Parallelization exploits the fact that each element can be updated independently (within one step): the grid can be divided up and the algorithm performed on each subgrid in separate processes or threads. The key is that neighbouring processes must exchange their subgrid boundary values as they are updated. We illustrate a one-dimensional decomposition of a square grid into three non-overlapping subgrids for three separate processes. Two Workers are allocated the end subgrids; the Master has the central subgrid, and controls the termination condition for all three processes. In addition to their allocated subgrid, each process maintains a copy of the boundary values ({\em ghost points}) of its neighbours; the new values are communicated after each iteration. This scheme allows the original grid to be divided in subgrids of any size. The session type between the Master and each of the two Workers from the side of the former is: \vspace{1mm} \noindent \begin{lstlisting} protocol masterToWorker { cbegin. // Request the Worker service. !<int>. // Send the size of the matrix. ![ // Enter the main loop (check termination condition). !<double[]>.?(double[]). /* Send our boundary values and.. ..get the Worker's updated ghost points. */ ?(double).?(double) // Receive the convergence data for Worker's subgrid. ]*. // After the last iteration.. ?(double[][]) // ..get the final results. } \end{lstlisting} \vspace{1mm} To control all the Workers simultaneously, the implementation of Master uses the SJ session constructs for multicasting output operations such as message-send and also session-iteration (see Appendix~\ref{app} for the full implementation). For example: \vspace{1mm} \noindent \hspace{1mm} \begin{tabular}{lp{4mm}l} \begin{lstlisting} // Master controls iteration condition. <mw1, mw2>.outwhile( // ![.. !accurateEnough(...) && iters < MAX_ITERS) { ... // Main body of the algorithm. } // ..]* \end{lstlisting} && \begin{lstlisting} // Workers obey the Master. <wm>.inwhile() { // ?[.. ... /* Main body of the algorithm. */ } // ..]* \end{lstlisting} \end{tabular} \\ \vspace{1mm} \noindent Like the standard while-statement, the outwhile operation evaluates the boolean condition for iteration (\SJ{!accurateEnough(...)} \SJ{\&\& iters $<$ MAX\_ITERS}) to determine whether the loop continues or terminates. The key difference is that this decision is implicitly communicated to the session peer (in this case, from Master to the two Worker), synchronising the control flow between two parties. Worker is programmed with the dual behaviour: \SJ{inwhile} does not specify a loop-condition because this decision is made by Master and communicated to Worker at each iteration. Inter-thread communication of large messages, such as arrays, can be optimised using SJ \SJ{noalias} types. A \SJ{noalias} variable on the RHS of an assignment or as a method argument --- such as to the \SJ{send} operation --- becomes \SJ{null} after the assignment or the method call. Combined with static type checking that precludes any potential assignment of aliased values to \SJ{noalias} targets, a \SJ{noalias} variable is guaranteed the sole reference to the pointed object at all times, permitting zero-copy message passing of \SJ{noalias} messages over compatible shared memory transports. In the present example, the \SJ{noalias} optimisation can be used to communicate the ghost point data; for example, the Worker implementations contain the following code extract. \vspace{1mm} \noindent \begin{lstlisting} // noalias array containing our boundary values (ghost points for the Master). noalias double[] ghostPoints = ...; /* Update and prepare our boundary values for sending. */ s_wm.send(ghostPoints); // Type-directed zero-copy send: !<noalias double[]> ... // ghostPoints variable becomes null. \end{lstlisting} \vspace{1mm} \noindent Transports that do not support this feature (e.g. TCP) can fall back to copy-on-send; the overall semantics of the program remains unchanged. This illustrates the {\em transport-independent} nature of SJ programs: the virtualisation of communication due to the SJ Runtime allows programs to make the best use of the whichever transports are available, {\em without} requiring any modification to the programs themselves. If the Master and Worker processes are run on separate machines, then the SJ Runtime can arrange, e.g. a TCP-based session; for the same programs, run as co-located threads, shared memory will be used. This SJ feature is further demonstrated for the next algorithm. \section{The $n$-Body Problem} \label{sec:nbody} The $n$-Body Problem involves finding the motion, according to classical mechanics, of a system of bodies given their masses and initial position and velocities. This advanced example demonstrates (1) the expressiveness of SJ and the extensions for complex iteration structures, by implementing an intricate circular communication pipeline; (2) SJ transport-independence (see \S~\ref{sec:bmarks}); and (3) the benefits of high-level message types (see \S~\ref{sec:conc}). Parallelism is achieved by dividing the particle set, and hence the calculations to determine the resultant force exerted on each body, amongst a collection of parallel processes. We use the approach where the processes, maintaining only the current state of their individual particle sets, are deployed to form a circular pipeline (ring topology). Firstly, the number of processes in the pipeline, $p$, is dynamically determined by sending a token around the ring. Then each step of the simulation involves $p-1$ iterations. In the first iteration, each process sends their particle data to their neighbour on the right and calculates the partial resultant forces exerted within their own particle set. In the n-th iteration, each process forwards on the particle data received in the previous iteration (line~({\em i}) in Figure~\ref{fig:mpi-sj2}), adds this data to the running force calculation ({\em ii}), and receives the next data set ({\em iii}). The particle data from the right neighbour is received by the end of the final iteration: each data set has now been seen by all processors in the pipeline, allowing the final results for the current simulation step to be calculated. The SJ implementation of the above algorithm has each process, i.e. each Worker unit in the pipeline, open a session server socket to accept a connection from its left neighbour, and create the connection to its right neighbour using a session client socket. The session type for the interaction in this algorithm, from the server side of each unit, is: \vspace{1mm} \noindent \hspace{1mm} \begin{tabular}{ll} \begin{lstlisting} protocol serverSide { sbegin. !<int>. ?[ ?[ ?(Particle[]) ]* ]* } \end{lstlisting} & \begin{lstlisting} // Interaction with the left neighbour. // Accept connection from left neighbour. // Forward on the ring initialisation token. // Main simulation loop (iteration flag received from the left). // Inner iterations within each simulation step. // Particle data forwarded through pipeline. (*@@*) (*@@*) (*@@*) \end{lstlisting} \end{tabular} \\ \vspace{1mm} \noindent The session type for the corresponding client side of each unit is simply the direct dual of \SJ{serverSide}: \SJ{protocol clientSide \{ cbegin.?(int).![![!<Particle[]>]*]*} \}, given by inverting the input (\SJ{?}) and output (\SJ{!}) symbols. For this client-server architecture, the ring topology is bootstrapped by designating two neighbouring processes to be the ``first'' and ``last'' pipeline units. The remaining SJ code for this example and a comparison with an MPI implementation (Figure~\ref{fig:mpi-sj2}) are outlined in \S~\ref{sec:conc}. \section{Performance Benchmarks} \label{sec:bmarks} This section presents performance measurements for the three parallel algorithms described above. The first two benchmarks show that the SJ Runtime, although still at an early implementation version with much scope for further optimisation, can perform competitively with MPJ Express \cite{mpjexpress}. Unlike Java MPI implementations built around JNI wrappers to C functions, MPJ Express adopts a pure Java approach which makes for a more informative comparison with SJ. The same machines in the same network environment were used for all the following benchmark experiments. Each machine is a dual-core Intel Core 2 Duo (Conroe~B2) at 2.13GHz with 2MB cache, 2GB main memory, running Ubuntu Linux~4.2.3 (kernel~2.6.24); the machines were connected via gigabit Ethernet, and the latency between two machines was measured using ping (64~Bytes) to be on average 0.10ms. The benchmark applications were compiled and executed using the standard Sun Java SE compiler and runtime versions~1.6.0. For each experiment, the results from 100 executions for each parameter configuration were recorded; here, we give the mean values. The full source code for the benchmark applications and the complete results can be found at \cite{parallelalgowebsite}. \paragraph{Monte Carlo $\pi$ approximation.} The first benchmark uses the SJ implementation of this algorithm to (1) verify the performance gain from increased parallelism, and (2) to compare the performance of the SJ Runtime against MPJ Express. Each process (Master, Workers and Client) was run on a separate machine, communicating via TCP. The results (Figure~\ref{fig:benchmarks12}), comparing both sequential and parallel versions of the algorithm, show that for a constant sample size (total number of test points), increasing the number of Workers indeed reduces the time to complete the algorithm proportionally. The results for the SJ implementation are around 5--6\% faster than the MPJ Express implementation \begin{figure}[t] \centering \begin{tabular}{|l||rr|} \hline \textbf{Configuration} & SJ (ms) & MPJ (ms) \\ \hline Sequential (1 Worker) & \multicolumn{2}{c|}{6717} \\ \hline 1 Master \& 1 Worker & 3764 & 3846 \\ \hline 1 Master \& 2 Workers & 2466 & 2606 \\ \hline 1 Master \& 3 Workers & 1885 & 1966 \\ \hline 1 Master \& 4 Workers & 1487 & 1579\\ \hline \end{tabular} \caption{ Monte Carlo $\pi$ for a varying number of Workers.} \label{fig:benchmarks12} \rule{\linewidth}{0.2mm} \end{figure} \paragraph{Jacobi Poisson solution.} The second benchmark, through the SJ implementation of the Jacobi iteration algorithm, demonstrates (1) the effectiveness of \SJ{noalias} types for zero-copy message transfer in a shared memory environment, and (2) again compares SJ performance to MPJ Express. Firstly, ``Ordinary'' (i.e. without \SJ{noalias}) and \SJ{noalias} versions of the Master and two Workers were run as co-VM threads on a single machine; the Client is connected to the Master from a separate machine via a TCP-session. We measured the time to complete the algorithm for square matrices of size (i.e. the length of one side of the matrix) 100 and 300. In both cases, the \SJ{noalias} version is approximately 20\% faster than the ordinary one (Figure~\ref{fig:benchmarks3}). For sizes greater than 300, we observed that the local computation costs start to dominate the communication costs for this fixed number of Workers, reducing the differences between the execution times of the ``Ordinary" and \SJ{noalias} versions, e.g. for matrix size 1000. Secondly, the distributed SJ implementation of Jacobi (the Client, Master and Workers run on separate machines connected via TCP) performs better than the MPJ Express implementation by 6\% on average (Figure~\ref{fig:benchmarks4}).\\ \begin{figure}[t] \centering \vspace{1mm} \subfigure[] { \begin{tabular}{|l||rr|} \hline \textbf{Matrix Size} & ``Ordinary'' (ms) & \SJ{noalias} (ms) \\ \hline 100 & 1270 & 992 \\ \hline 300 & 24436 & 19448 \\ \hline 1000 & 288532 & 299279 \\ \hline \end{tabular} \label{fig:benchmarks3} } \hspace{5mm} \subfigure[]{ \begin{tabular}{|l||rr|} \hline \textbf{Matrix Size} & SJ (ms) & MPJ (ms) \\ \hline 100 & 3713 & 4460 \\ \hline 300 & 19501 & 19834\\ \hline \end{tabular} \label{fig:benchmarks4} } \vspace{-2mm} \caption{(a) Jacobi: ``ordinary'' vs. \SJ{noalias} versions; (b) Jacobi: SJ vs. MPJ Express.} \label{fig:benchmark3} \rule{\linewidth}{0.2mm} \end{figure} \paragraph{$n$-Body simulation.} The third benchmark uses the $n$-Body simulation to demonstrate the important improvement in productivity enabled by SJ transport-independence: this single SJ implementation was run in the different communication environments (locally concurrent, distributed), making the best use of the available transports (TCP, shared memory, etc.), without {\em any} changes to the source code for the Workers (although the shared memory version required a few lines of external code to bootstrap the Workers as Java threads). The benchmark was executed using two pipeline Worker units (not using \SJ{noalias}) in three different configurations: the two Workers on separate machines using TCP (Distributed), as separate processes on the same machine using TCP (Localhost), and as co-VM threads using shared memory (Threads). We recorded the results for simulations involving 100, 300 and 1000 particles, distributed equally between the Workers. As expected, the results (Figure~\ref{fig:benchmark4}) show the Threads version is faster than Localhost: around 27\% for 100 particles, 24\% for 300, and 10\% for 1000. The Distributed version is in turn slightly slower (latency is very low) than Localhost: 10\% for 100 particles, 4\% for 300, and 3\% for 1000. The relative performance gain between each version decreases for larger particle sets because the local computation costs begin to dominate the communication costs for this fixed number of Workers. Naturally, performance can be improved for simulations involving many particles by increasing the degree of parallelism, i.e. using more Workers.\\ \begin{figure}[t] \centering \begin{tabular}{|l| |rrr|} \hline \textbf{Particles} & Distrib. (ms) & Localhost (ms) & Threads (ms) \\ \hline 100 & 496 & 452 & 326\\ \hline 300 & 1194 & 1144 & 865 \\ \hline 1000 & 7702 & 7497 & 6785\\ \hline \end{tabular} \caption{$n$-Body simulation: Distributed vs. Localhost vs. Threads versions.} \label{fig:benchmark4} \rule{\linewidth}{0.2mm} \end{figure} \section{SJ and MPI Comparison} \label{sec:conc} This section compares SJ against MPI in terms of language support for communications programming, with reference to MPI implementations of the above algorithms \cite{usingmpi}. Since MPI has an extensive library of functions developed over 15 years, many of these are not yet directly supported in SJ, e.g. MPI Jacobi makes use of a virtual topology (\SJ{MPI\_Cart\_Create}) and collective data movement operations (\SJ{MPI\_Bcast} and \SJ{MPI\_Allreduce}, for broadcasting the matrix size and distributing the termination condition in (2)). However, many of these features can be encoded into a session type, as shown above. Furthermore, we observed the following benefits of SJ against MPI. \paragraph{Type and communication safety from session types.} MPI is designed as a portable API specification to be implemented for varying host languages. Coupled to the low-level nature of many MPI functions, the design of accompanying MPI program verification techniques for a host language can be difficult. Common MPI errors recognized by the community include: \begin{itemize} \setlength{\topsep}{0mm} \setlength{\leftmargin}{0mm} \setlength{\itemindent}{0mm} \setlength{\itemsep}{0mm} \setlength{\listparindent}{0mm} \item{{\bf Invalid actions before \SJ{MPI\_Init} and after \SJ{MPI\_Finalize}}.} The execution of such MPI operations can lead to runtime errors such as broken invariants, messages not broadcasted, and incorrect collective operations. Figure~\ref{fig:mpi-sj} presents the correct code of setting up the topology in the $n$-body simulation in MPI\footnote{This MPI implementation of the $n$-Body simulation is taken from the Using MPI website \cite{MPI}.} (left column) and SJ (right column). In the MPI code, the errors we are referring to would come from adding MPI operations before line~\ref{line:mpi1} and after line~\ref{line:mpi2}. In SJ, actions incorrectly performed before the server socket (line~\ref{line:sj1}) or the session (lines~\ref{line:sj2}--\ref{line:sj3}) have been initialised are rejected by the compiler. The static type system of SJ also does not allow session actions to be performed after leaving the relevant session-try scope (i.e. on \SJ{left} or \SJ{right} after line~\ref{line:sj4}). The MPI and SJ code for the main body of the algorithm is given in Figure~\ref{fig:mpi-sj2}. \item{{\bf Unmatched \SJ{MPI\_Send} and \SJ{MPI\_Recv}}.} Such errors can lead to a mismatch between the sent and expected message type/structure, or a variety of deadlock situations depending on the communication mode. For example, two processes deadlock if each is waiting for a message before sending the message expected by the other. In the standard (buffer-blocking) mode, the converse situation (both processes attempting to send before receiving) can also deadlock: if both message sizes are bigger than the available space in the medium and opposing receive buffers, then the processes cannot complete their write operations. A related problem is matching a \SJ{MPI\_Bcast} output with \SJ{MPI\_Recv}. Standard usage is to receive a broadcast message using the complementary \SJ{MPI\_Bcast} input. \SJ{MPI\_Recv} consumes the message; hence, the receiver must be able to determine which processes have not yet seen the message and manually re-broadcast it. \item{{\bf Concurrency issues}.} Incorrect access of a shared communicator by separate threads can violate the intended message causalities between the sender(s) and the receivers. In addition, race conditions can arise due to modifying, or even just by accessing, messages that are in transit. \end{itemize} \begin{figure}[t] {\lstset{ basicstyle=\small, numbers=left, numberstyle=\tiny, stepnumber=1, numbersep=5pt } \noindent \begin{tabular}{ll} \begin{lstlisting} main(int argc, char *argv[]) { // Set up of the topology. MPI_Init(&argc, &argv); (*@\label{line:mpi1}@*) MPI_Comm_rank(MPI_COMM_WORLD, &rank); MPI_Comm_size(MPI_COMM_WORLD, &size); // Get the best ring in the topology. periodic = 1; MPI_Cart_create(MPI_COMM_WORLD, 1, &size, &periodic, 1, &commring); MPI_Cart_shift(commring, 0, 1, &left, &right); ... // Main algorithm body. MPI_Finalize(); (*@\label{line:mpi2}@*) return 0; } (*@~@*) (*@~@*) (*@~@*) \end{lstlisting} & \begin{lstlisting} public void run(...) { // Set up the sockets for the topology. SJService c_r = SJService.create(pc_nbody, host_r, port_r); SJServerSocket ss_l; SJSocket left, right; try(ss_l) { ss_l = SJServerSocket.create(ps_nbody, port_l); (*@\label{line:sj1}@*) try(left, right) { left = ss_l.accept(); (*@\label{line:sj2}@*) right = c_r.request(); (*@\label{line:sj3}@*) // Determine the topology size. left.send(right.receiveInt() + 1); ... // Main algorithm body. } finally {...} (*@\label{line:sj4}@*) } catch(SJIncompatibleSessionException ise) {...} ...// Handling for other exceptions. } \end{lstlisting} \end{tabular}} \caption{Setting up the topology for the $n$-Body simulation in MPI and in SJ.} \label{fig:mpi-sj} \rule{\linewidth}{0.2mm} \end{figure} \noindent As illustrated in the previous sections, {\bf\em SJ programs are guaranteed free from all of the above errors} by the semantics of session communication and static session type checking. The first two points are directly prevented by the properties of session types. For the third point, the SJ compiler disallows sharing of session socket objects (implicitly \SJ{noalias}), and message copying/linear transfer can be safely and explicitly controlled via \SJ{noalias} types. \paragraph*{High-level message types.} In many parallel algorithms, messages are mainly communicated via arrays. For MPI, effort is required to manually track and communicate array indices, e.g.~for message length or the number of messages. In contrast, the high-level type-abstraction for messages allows SJ programmers to treat both object and primitive array messages as regular Java array objects. For instance, the MPI version of the main algorithm for the $n$-Body simulation\footnotemark[1] (Figure~\ref{fig:mpi-sj2}, left) broadcasts the number of particles managed by each process, through the \SJ{MPI_Allgather} operation (line~\ref{line:mpi21}). Thus, the amount of data to be read from each particle set (line~\ref{line:mpi22}) can be determined (lines~\ref{line:mpi23}--\ref{line:mpi24}). In SJ (Figure~\ref{fig:mpi-sj2}, right), the particle data is simply received as discrete array messages (line~\ref{line:sj21}), avoiding manual handling of message sizes. Therefore, the MPI code between lines~\ref{line:mpi21}--\ref{line:mpi24} is unnecessary in the SJ implementation. The rest of the code structure is the same in both implementations. In the SJ implementation of the $n$-Body, the assignment in (\emph{iii}) is permitted because the received message is implicitly \SJ{noalias}. \begin{figure}[t!] {\lstset{ basicstyle=\small, numbers=left, numberstyle=\tiny, stepnumber=1, numbersep=5pt } \noindent \begin{tabular}{ll} \begin{lstlisting} // Get the sizes and displacements. MPI_Allgather(&npart, 1, MPI_INT, counts, (*@\label{line:mpi21}@*) 1, MPI_INT, commring); displs[0] = 0; (*@\label{line:mpi23}@*) for(i=1; i<size; i++) displs[i] = displs[i-1] + counts[i-1]; totpart = displs[size-1] + counts[size-1]; (*@\label{line:mpi24}@*) InitParticles(particles, pv, npart); while(cnt--) { double max_f, max_f_seg; // Load the initial sendbuffer. memcpy(sendbuf, particles, npart * sizeof(Particle)); for(pipe=0; pipe<size; pipe++) { if(pipe != size-1) { MPI_Isend(sendbuf, npart, particletype, right, pipe, commring, &request[0]); MPI_Irecv(recvbuf, npart, particletype, (*@\label{line:mpi22}@*) left, pipe, commring, &request[1]); } // Compute forces. max_f_seg = ComputeForces(particles, sendbuf, pv, npart); // Wait for non-blocking receives to return. if(pipe != size-1) MPI_Waitall(2, request, statuses); memcpy(sendbuf, recvbuf, counts[pipe] * sizeof(Particle)); } // Update our own particle data. sim_t += ComputeNewPos(particles, pv, npart, max_f, commring); } \end{lstlisting} &\quad \begin{lstlisting} initParticles(particles, pvs); /* Synchronise with our two neighbours for each simulation step. */ right.outwhile(left.inwhile()) { // Load the initial sendbuffer. Particle[] current = new Particle[numParticles]; System.arraycopy(particles, 0, current, 0, numParticles); /* Inner iterations within each simulation step. */ right.outwhile(left.inwhile()) { // (*@(\emph{i})@*) Forward the current data set. right.send(current); /* (*@(\emph{ii})@*) Add the current data to the running calculation. */ computeForces(particles, current, pvs); // (*@(\emph{iii})@*) Receive the next data set. current = (Particle[]) left.receive(); (*@\label{line:sj21}@*) } /* Calculate the final results for this simulation step and update our own particle data. */ computeForces(particles, current, pvs); computeNewPos(particles, pvs, i); i++; } \end{lstlisting} \end{tabular}} \caption{Implementing the main body of the $n$-Body simulation algorithm in MPI and SJ.} \label{fig:mpi-sj2} \rule{\linewidth}{0.2mm} \end{figure} \paragraph{Transparent zero-copy message passing.} SJ provides direct language support for zero-copy transfer in shared memory contexts through \SJ{noalias} types. This feature can enable significant performance increases for multi-threaded programs (see \S~\ref{sec:bmarks}). Moreover, the communication of \SJ{noalias} types retains consistent semantics in all transport contexts (see {\em transport-independence} in \S~\ref{sec:jacobi}). \section{Conclusions and Future Work} We demonstrated expressiveness, productivity and performance benefits of session-based programming in SJ through the presented parallel algorithm implementations. Although we have seen that the above algorithms were readily implemented in the current SJ, immediate future work includes expanding the set of SJ operations and constructs, e.g. with session typed equivalents of MPI functions and features that are not yet directly supported. For example, whilst the MPI {\em standard} mode (send and receive block on their respective buffers) corresponds to the session communication semantics in SJ, MPI has several additional modes: {\em synchronous} (send and receive operations synchronise), {\em ready} (programmer notifies the system that a receive has been posted), and {\em buffered} (user manually handles send buffers). We also wish to compare SJ to PGAS languages such as X10 \cite{X10Homepage} using parallel algorithm implementation as a basis. We believe that extending SJ with full multiparty session types \cite{DBLP:conf/popl/HondaYC08} would allow richer topologies such as the ring and 2D-mesh to be expressed more naturally in a type-safe manner. For example, the SJ $n$-Body implementation currently requires creating one intermediary session (for the final pipeline link) in each simulation step; with multiparty sessions, we would only need to open a single session for the complete simulation. Our prediction is that multiparty sessions will offer better support for massive parallelism than the current client-server based session sockets. We plan to identify design issues and possible overheads for global type-checking through further implementation of parallel algorithms with complex communication patterns. SJ programs are guaranteed free from type and communication errors, and perform competitively against other Java communication runtimes. In certain cases, SJ programs can out-perform their counterparts implemented in communication-safe systems such as RMI \cite{SJecoop08} and also lower-level, non communication-safe message passing systems such as MPJ Express (\S~\ref{sec:bmarks}). \section{Acknowledgments} \label{sec:ack} We thank the reviewers for their helpful comments on the submission. We also thank Kohei Honda and Vijay Saraswat for their comments on a first draft of this paper. The work is partially supported by EPSRC GR/T03208 and GR/T03215. \bibliographystyle{eptcs}
\section{Introduction} \label{sec:intro} The term ``{\bf flavors}'' is used, in the jargon of particle physics, to describe several copies of the same gauge representation, namely several fields that are assigned the same quantum charges. Within the Standard Model (SM), when thinking of its unbroken $SU(3)_{\rm C}\times U(1)_{\rm EM}$ gauge group, there are four different types of particles, each coming in three flavors: \begin{itemize} \item Up-type quarks in the $(3)_{+2/3}$ representation: $u,c,t$; \item Down-type quarks in the $(3)_{-1/3}$ representation: $d,s,b$; \item Charged leptons in the $(1)_{-1}$ representation: $e,\mu,\tau$; \item Neutrinos in the $(1)_{0}$ representation: $\nu_1,\nu_2,\nu_3$. \end{itemize} The term ``{\bf flavor physics}'' refers to interactions that distinguish between flavors. By definition, gauge interactions, namely interactions that are related to unbroken symmetries and mediated therefore by massless gauge bosons, do not distinguish among the flavors and do not constitute part of flavor physics. Within the Standard Model, flavor-physics refers to the weak and Yukawa interactions. With New Physics (NP), there are likely to be additional `flavored' interactions. The term ``{\bf flavor changing}'' refers to processes where the initial and final flavor-numbers (that is, the number of particles of a certain flavor minus the number of anti-particles of the same flavor) are different. In ``flavor changing charged current'' processes, both up-type and down-type flavors, and/or both charged lepton and neutrino flavors are involved. Examples are (i) muon decay via $\mu\to e\bar\nu_i\nu_j$, and (ii) $K^-\to\mu^-\bar\nu_j$ (which corresponds, at the quark level, to $s\bar u\to\mu^-\bar\nu_j$). Within the Standard Model, these processes are mediated by the $W$-bosons and occur at tree level. In ``{\bf flavor changing neutral current}'' (FCNC) processes, either up-type or down-type flavors but not both, and/or either charged lepton or neutrino flavors but not both, are involved. Example are (i) muon decay via $\mu\to e\gamma$ and (ii) $K_L\to\mu^+\mu^-$ (which corresponds, at the quark level, to $s\bar d\to\mu^+\mu^-$). Within the Standard Model, these processes do not occur at tree level, and are often highly suppressed. This situation makes FCNC particularly sensitive to new physics: If the new physics does not have the same flavor suppression factors as the standard model, then it could contribute to FCNC comparably to the standard model even if it takes places at energy scales that are orders of magnitude higher than the weak scale. The fact that flavor physics is a very sensitive probe of high energy physics is the main reason for the experimental effort to measure flavor parameters and the theoretical effort to interpret these data. \section{The Standard Model} \label{sec:sm} The Standard Model (SM) is defined as follows:\\ (i) The gauge symmetry is \begin{equation}\label{smsym} G_{\rm SM}=SU(3)_{\rm C}\times SU(2)_{\rm L}\times U(1)_{\rm Y}. \end{equation} It is spontaneously broken by the vacuum expectation values (VEV) of a single Higgs scalar, $\phi(1,2)_{1/2}$ ($\langle\phi^0\rangle=v/\sqrt{2}$): \begin{equation}\label{smssb} G_{\rm SM} \to SU(3)_{\rm C}\times U(1)_{\rm EM}. \end{equation} (ii) There are three fermion generations, each consisting of five representations of $G_{\rm SM}$: \begin{equation}\label{ferrep} Q_{Li}(3,2)_{+1/6},\ \ U_{Ri}(3,1)_{+2/3},\ \ D_{Ri}(3,1)_{-1/3},\ \ L_{Li}(1,2)_{-1/2},\ \ E_{Ri}(1,1)_{-1}. \end{equation} The Standard Model Lagrangian, ${\cal L}_{\rm SM}$, is the most general renormalizable Lagrangian that is consistent with the gauge symmetry (\ref{smsym}), the particle content (\ref{ferrep}) and the pattern of spontaneous symmetry breaking (\ref{smssb}). It can be divided to three parts: \begin{equation}\label{LagSM} {\cal L}_{\rm SM}={\cal L}_{\rm kinetic+gauge}+{\cal L}_{\rm Higgs} +{\cal L}_{\rm Yukawa}. \end{equation} The source of all flavor physics is in the Yukawa interactions: \begin{equation}\label{Hqint} -{\cal L}_{\rm Yukawa}=Y^d_{ij}~{\overline{Q}_{Li}}\phi D_{Rj} +Y^u_{ij}~{\overline{Q}_{Li}}\tilde\phi U_{Rj}+Y^e_{ij}~{\overline{L}_{Li}}\phi E_{Rj} +{\rm h.c.}. \end{equation} (where $\tilde\phi=i\tau_2\phi^\dagger$). This part of the Lagrangian is, in general, flavor-dependent (that is, $Y^f\not\propto{\bf 1}$) and CP violating. In the absence of the Yukawa matrices $Y^d$, $Y^u$ and $Y^e$, the SM has a large $U(3)^5$ global symmetry: \begin{equation}\label{gglobal} G_{\rm global}(Y^{u,d,e}=0)=SU(3)_q^3\times SU(3)_\ell^2\times U(1)^5. \end{equation} The non-Abelian part of this symmetry is particularly relevant to flavor physics: \begin{eqnarray}\label{susuu} SU(3)_q^3&=&SU(3)_Q\times SU(3)_U\times SU(3)_D,\nonumber\\ SU(3)_\ell^2&=&SU(3)_L\times SU(3)_E. \end{eqnarray} The point that is important for our purposes is that ${\cal L}_{\rm kinetic+gauge}+{\cal L}_{\rm Higgs}$ respect the non-Abelian flavor symmetry $S(3)_q^3\times SU(3)_\ell^2$, under which \begin{equation}\label{symkh} Q_L\to V_QQ_L,\ \ \ U_R\to V_U U_R,\ \ \ D_R\to V_D D_R,\ \ L_L\to V_L L_L,\ \ \ E_R\to V_E E_R, \end{equation} where the $V_i$ are unitary matrices. The Yukawa interactions (\ref{Hqint}) break the global symmetry, \begin{equation}\label{globre} G_{\rm global}(Y^{u,d,e}\neq0)= U(1)_B\times U(1)_e\times U(1)_\mu\times U(1)_\tau. \end{equation} (Of course, the gauged $U(1)_Y$ also remains a good symmetry.) Thus, the transformations of Eq. (\ref{symkh}) are not a symmetry of ${\cal L}_{\rm SM}$. Instead, they correspond to a change of the interaction basis. These observations also offer an alternative way of defining flavor physics: it refers to interactions that break the $SU(3)^5$ symmetry (\ref{symkh}). Thus, the term ``{\bf flavor violation}'' is often used to describe processes or parameters that break the symmetry. Using the transformation (\ref{symkh}), one can choose an interaction basis where the number of parameters is minimized. A useful example of such a basis for the quark Yukawa matrices is the following: \begin{equation}\label{speint} Y^d=\lambda_d,\ \ \ Y^u=V^\dagger\lambda_u, \end{equation} where $\lambda_{d,u}$ are diagonal, \begin{equation}\label{deflamd} \lambda_d={\rm diag}(y_d,y_s,y_b),\ \ \ \lambda_u={\rm diag}(y_u,y_c,y_t), \end{equation} while $V$ is a unitary matrix that depends on three real angles and one complex phase. We conclude that there are 10 quark flavor parameters: 9 real ones and a single phase. In the mass basis, one identifies six of the real parameters as the six quark masses, while the remaining three real and one imaginary parameters appear in the CKM matrix \cite{Cabibbo:1963yz,Kobayashi:1973fv}, $V$, which describes the couplings of the charged weak-force carriers, the $W^\pm$-bosons, with quark-antiquark pairs. Within the standard model, all flavor changing physics comes from the $V$ matrix. There are various ways to choose the four parameters of $V$. One of the most convenient ways is given by the Wolfenstein parametrization, where the four mixing parameters are $(\lambda,A,\rho,\eta)$ with $\lambda=|V_{us}|=0.23$ playing the role of an expansion parameter and $\eta$ representing the CP violating phase \cite{Wolfenstein:1983yz,Buras:1994ec}: \begin{equation}\label{wolpar} V\simeq\left(\begin{matrix} 1-\frac12\lambda^2-\frac18\lambda^4 & \lambda & A\lambda^3(\rho-i\eta)\cr -\lambda +\frac12A^2\lambda^5[1-2(\rho+i\eta)] & 1-\frac12\lambda^2-\frac18\lambda^4(1+4A^2) & A\lambda^2 \cr A\lambda^3[1-(1-\frac12\lambda^2)(\rho+i\eta)]& -A\lambda^2+\frac12A\lambda^4[1-2(\rho+i\eta)] & 1-\frac12A^2\lambda^4 \cr\end{matrix}\right) . \end{equation} One can assume that flavor changing processes are fully described by the SM, and check the consistency of the various measurements with this assumption. The values of $\lambda$ and $A$ are known rather accurately \cite{Amsler:2008zzb} from, respectively, $K\to\pi\ell\nu$ and $b\to c\ell\nu$ decays: \begin{equation}\label{lamaexp} \lambda=0.2257\pm0.0010,\ \ \ A=0.814\pm0.022. \end{equation} Then, one can express all the relevant observables as a function of the two remaining parameters, $\rho$ and $\eta$, and check whether there is a range in the $\rho-\eta$ plane that is consistent with all measurements. The list of observables includes the following: \begin{itemize} \item The rates of inclusive and exclusive charmless semileptonic $B$ decays depend on $|V_{ub}|^2\propto\rho^2+\eta^2$; \item The CP asymmetry in $B\to\psi K_S$, $S_{B\to \psi K}=\sin2\beta=\frac{2\eta(1-\rho)}{(1-\rho)^2+\eta^2}$; \item The rates of various $B\to DK$ decays depend on the phase $\gamma$, where $e^{i\gamma}=\frac{\rho+i\eta}{\sqrt{\rho^2+\eta^2}}$; \item The rates of various $B\to\pi\pi,\rho\pi,\rho\rho$ decays depend on the phase $\alpha=\pi-\beta-\gamma$; \item The ratio between the mass splittings in the neutral $B$ and $B_s$ systems is sensitive to $|V_{td}/V_{ts}|^2=\lambda^2[(1-\rho)^2+\eta^2]$; \item The CP violation in $K\to\pi\pi$ decays, $\epsilon_K$, depends in a complicated way on $\rho$ and $\eta$. \end{itemize} The resulting constraints are shown in Fig. \ref{fg:UT}. \begin{figure}[tb] \centering {\includegraphics[width=0.65\textwidth]{utfit09.eps}} \caption{Allowed region in the $\rho,\eta$ plane. Superimposed are the individual constraints from charmless semileptonic $B$ decays ($|V_{ub}/V_{cb}|$), mass differences in the $B^0$ ($\Delta m_d$) and $B_s$ ($\Delta m_s$) neutral meson systems, and CP violation in $K\to\pi\pi$ ($\varepsilon_K$), $B\to\psi K$ ($\sin2\beta$), $B\to\pi\pi,\rho\pi,\rho\rho$ ($\alpha$), and $B\to DK$ ($\gamma$). Taken from \cite{ckmfitter}.} \label{fg:UT} \end{figure} The consistency of the various constraints is impressive. In particular, the following ranges for $\rho$ and $\eta$ can account for all the measurements \cite{Amsler:2008zzb}: \begin{equation} \rho=+0.135^{+0.031}_{-0.016},\ \ \ \eta=+0.349\pm0.017. \end{equation} One can make then the following statement \cite{Nir:2002gu,Ligeti:2004ak,NMFV}:\\ {\bf Very likely, flavor violation and CP violation in flavor changing processes are dominated by the CKM mechanism.} One can actually go a step further, and allow for arbitrary new physics in all flavor changing processes except for those that have contributions from SM tree diagrams. Then, one can quantitatively constrain the size of new physics contributions to processes such as neutral meson mixing. We do so in the next section. \section{Model Independent Constraints} \label{sec:indep} In order to describe NP effects in flavor physics we can follow two main strategies: (i) build an explicit ultraviolet completion of the model, and specify which are the new fields beyond the SM ones, or (ii) analyse the NP effects using a generic effective-theory approach, or integrating-out the new heavy fields. The first approach is more predictive, but also more model dependent. We follow this approach in Sect.~\ref{sec:susy} and~\ref{sec:exdim} in two well-motivated SM extensions. In this and the next section we follow the second strategy, which is less predictive but also more general. Assuming the new degrees to be heavier than SM fields, we can integrate them out and describe NP effects by means of a generalization of the Fermi Theory. The SM Lagrangian becomes the renormalizable part of a more general local Lagrangian which includes an infinite tower of operators with dimension $d>4$, constructed in terms of SM fields, suppressed by inverse powers of an effective scale $\Lambda > M_W$: \begin{equation} \mathcal{L}_{\rm eff} = \mathcal{L}_{\rm SM} + \sum ~ \frac{c_{i}^{(d)}}{\Lambda^{(d-4)}} ~ O_i^{(d)}({\rm SM~fields}). \label{eq:effL} \end{equation} This general bottom-up approach allows us to analyse all realistic extensions of the SM in terms of a limited number of parameters (the coefficients of the higher-dimensional operators). The drawback of this method is the impossibility to establish correlations of NP effects at low and high energies: the scale $\Lambda$ defines the cut-off of the effective theory. However, correlations among different low-energy processes can still be established implementing specific symmetry properties, such as the MFV hypothesis (Sect.~\ref{sec:mfv}). The experimental tests of such correlations allow us to test/establish general features of the new theory which holds independently of the dynamical details of the model. In particular, $B$, $D$ and $K$ decays are extremely useful in determining the flavor-symmetry breaking pattern of the NP model. \subsection{Bounds from $\Delta F=2$ down-type transitions} The starting points for this analysis is the observation that in several realistic NP models we can neglect non-standard effects in all cases where the corresponding effective operator is generated at the tree-level within the SM. This general assumption implies that the experimental determination of the CKM matrix via tree-level processes is free from the contamination of NP contributions. Using this determination we can unambiguously predict meson-antimeson mixing and FCNC amplitudes within the SM and compare it with data, constraining the couplings of the $\Delta F=2$ operators in~(\ref{eq:effL}). Each $\Delta F=2$ amplitude is then conveniently parametrized in terms of the shift induced in the modulo and the CPV phase, or the real and imaginary part~\cite{Silva:1996ih,Grossman:1997dd}: \begin{equation} \frac{\langle B_q| \mathcal{L}_\mathrm{eff}| \overline{B}_q\rangle}{\langle B_q| \mathcal{L}_\mathrm{SM} | \overline{B}_q\rangle} =C_{B_q} e^{2 i \phi_{B_q}}~, \qquad \frac{{\rm Re}[\langle K^0| \mathcal{L}_\mathrm{eff}| \overline{K^0} \rangle]}{{\rm Re}[\langle K^0 | \mathcal{L}_\mathrm{SM} | \overline{K^0} \rangle]} =C_{\Delta m_K} ~\stackrel{{\rm Re} \to {\rm Im}}{\longrightarrow}~ C_{\epsilon_K}~, \end{equation} An updated analysis of these constraints has been presented in~\cite{Bona:2007vi} (see Fig.~\ref{fig:UTfit}). The main conclusions that can be drawn from this analysis can be summarized as follows: \begin{figure}[t] \begin{center} \includegraphics[width=70mm]{BB_UTfit.eps} \includegraphics[width=70mm]{KK_UTfit.eps} \caption{% Constraints on the effective parameters encoding NP effects in $B_d$--$\overline{B}_d$ mixing and $K^0$--$\overline{K^0}$ mixing as obtained by the UTfit collaboration~\cite{Bona:2007vi}.} \label{fig:UTfit} \end{center} \end{figure} (i) In all the three accessible short-distance amplitudes ($K^0$--$\overline{K^0}$, $B_d$--$\overline{B}_d$, and $B_s$--$\overline{B}_s$) the magnitude of the new-physics amplitude cannot exceed, in size, the SM short-distance contribution. The latter is suppressed by both the GIM mechanism and the hierarchical structure of the CKM matrix, \begin{equation} \mathcal{A}_{\rm SM}^{\Delta F=2} \approx \frac{ G_F^2 m_t^2 }{16 \pi^2} \left(V_{ti}^* V_{tj} \right)^2 \times \langle \overline{M} | (\overline{Q}_{Li} \gamma^\mu Q_{Lj} )^2 | M \rangle \times F\left(\frac{M_W^2}{m_t^2}\right), \end{equation} where $F$ is a loop function of $\mathcal{O}(1)$. As a result, new-physics models with TeV-scale flavored degrees of freedom and $\mathcal{O}(1)$ effective flavor-mixing couplings are ruled out. To set explicit bounds, let us consider for instance the subset of left-handed $\Delta F=2$ operators in the generic effective Lagrangian in (\ref{eq:effL}), namely \begin{equation}\label{eq:qlql} \Delta \mathcal{L}^{\Delta F=2} = \sum_{i\not=j} \frac{c_{ij}}{\Lambda^2} (\overline{Q}_{Li} \gamma^\mu Q_{Lj} )^2~, \end{equation} where the $c_{ij}$ are dimensionless couplings. The condition $|\mathcal{A}^{\Delta F=2}_{\rm NP}| < |\mathcal{A}^{\Delta F=2}_{\rm SM} |$ implies \begin{eqnarray} \Lambda > \frac{ 4.4~{\rm TeV} }{| V_{ti}^* V_{tj}|/|c_{ij}|^{1/2} } \sim \left\{ \begin{array}{l} 1.3\times 10^4~{\rm TeV} \times |c_{sd}|^{1/2} \!\!\!\!\!\!\! \\ 5.1\times 10^2~{\rm TeV} \times |c_{bd}|^{1/2} \!\!\!\!\!\!\! \\ 1.1\times 10^2~{\rm TeV} \times |c_{bs}|^{1/2} \!\!\!\!\!\!\! \end{array} \right. \label{eq:bound} \end{eqnarray} The strong bounds on $\Lambda$ for generic $c_{ij}$ of order 1 is a manifestation of what in many specific frameworks (supersymmetry, technicolor, etc.) goes under the name of {\em flavor problem}: if we insist that the new physics emerges in the TeV region, we have to conclude that it possesses a highly non-generic flavor structure. (ii) In the case of $B_d$--$\overline{B}_d$ and $K^0$--$\overline{K^0}$ mixing, where both CP conserving and CP-violating observables are measured with excellent accuracy, there is still room for a sizable NP contribution (relative to the SM one), provided that it is to a good extent aligned in phase with the SM amplitude [${\cal O}\left(0.01\right)$ for the $K$ system and ${\cal O}\left(0.3\right)$ for the $B_d$ system]. This is because the theoretical errors in the observables used to constraint the phases, $S_{B_d \to \psi K}$ and $\epsilon_K$, are smaller with respect to the theoretical uncertainties in $\Delta m_{B_d}$ and $\Delta m_K$, which constrain the magnitude of the mixing amplitudes. (iii) In the case of $B_s$--$\overline{B}_s$ mixing, the precise determination of $\Delta m_{B_s}$ does not allow large deviations in modulo with respect to the SM. The constraint is particularly severe if we consider the ratio $\Delta m_{B_d}/\Delta m_{B_s}$, where hadronic uncertainties cancel to a large extent. However, the constraint on the CP-violating phase is quite poor. Present data from CDF~\cite{Aaltonen:2007he} and D0~\cite{:2008fj} indicate a large central value for the CP-violating phase, contrary to the SM expectation. The errors are, however, still large and the disagreement with the SM is at about the $2 \sigma$ level. If the disagreement persists, becoming statistically significant, this would not only signal the presence of physics beyond the SM, but would also rule out a whole subclass of MFV models (see Sect.~\ref{sec:mfv}). (iv) In $D-\bar D$ mixing we cannot estimate the SM contribution from first principles; however, to a good accuracy this is CP conserving. As a result, strong bounds on possible non-standard CP-violating contributions can still be set. The resulting constraints are only second to those from $\epsilon_K$, and unlike in the case of $\epsilon_K$ are controlled by experimental statistics and could possibly be significantly improved in the near future. \begin{table}[t] \begin{center} \begin{tabular}{c|c c|c c|c} \hline\hline \rule{0pt}{1.2em}% Operator & \multicolumn{2}{c|}{Bounds on $\Lambda$~in~TeV~($c_{ij}=1$)} & \multicolumn{2}{c|}{Bounds on $c_{ij}$~($\Lambda=1$~TeV) }& Observables\cr & Re& Im & Re & Im \cr \hline $(\bar s_L \gamma^\mu d_L )^2$ &~$9.8 \times 10^{2}$& $1.6 \times 10^{4}$ &$9.0 \times 10^{-7}$& $3.4 \times 10^{-9}$ & $\Delta m_K$; $\epsilon_K$ \\ ($\bar s_R\, d_L)(\bar s_L d_R$) & $1.8 \times 10^{4}$& $3.2 \times 10^{5}$ &$6.9 \times 10^{-9}$& $2.6 \times 10^{-11}$ & $\Delta m_K$; $\epsilon_K$ \\ \hline $(\bar c_L \gamma^\mu u_L )^2$ &$1.2 \times 10^{3}$& $2.9 \times 10^{3}$ &$5.6 \times 10^{-7}$& $1.0 \times 10^{-7}$ & $\Delta m_D$; $|q/p|, \phi_D$ \\ ($\bar c_R\, u_L)(\bar c_L u_R$) & $6.2 \times 10^{3}$& $1.5 \times 10^{4}$ &$5.7 \times 10^{-8}$& $1.1 \times 10^{-8}$ & $\Delta m_D$; $|q/p|, \phi_D$\\ \hline$(\bar b_L \gamma^\mu d_L )^2$ & $5.1 \times 10^{2}$ & $9.3 \times 10^{2}$ & $3.3 \times 10^{-6}$ & $1.0 \times 10^{-6}$ & $\Delta m_{B_d}$; $S_{\psi K_S}$ \\ ($\bar b_R\, d_L)(\bar b_L d_R)$ & $1.9 \times 10^{3}$ & $3.6 \times 10^{3}$ & $5.6 \times 10^{-7}$ & $1.7 \times 10^{-7}$ & $\Delta m_{B_d}$; $S_{\psi K_S}$ \\ \hline $(\bar b_L \gamma^\mu s_L )^2$ & \multicolumn{2}{c|}{$1.1 \times 10^{2}$} & \multicolumn{2}{c|}{$7.6\times10^{-5}$} & $\Delta m_{B_s}$ \\ ($\bar b_R \,s_L)(\bar b_L s_R)$ & \multicolumn{2}{c|}{$3.7 \times 10^{2}$} & \multicolumn{2}{c|}{$1.3\times10^{-5}$} & $\Delta m_{B_s}$ \\ \hline\hline \end{tabular} \caption{\label{tab:DF2} Bounds on representative dimension-six $\Delta F=2$ operators. Bounds on $\Lambda$ are quoted assuming an effective coupling $1/\Lambda^2$, or, alternatively, the bounds on the respective $c_{ij}$'s assuming $\Lambda=1$ TeV. Observables related to CPV are separated from the CP conserving ones with semicolons. In the $B_s$ system we only quote a bound on the modulo of the NP amplitude derived from $\Delta m_{B_s}$ (see text). For the definition of the CPV observables in the $D$ system see Ref.~\cite{Bergmann:2000id}. } \end{center} \end{table} A more detailed list of the bounds derived from $\Delta F=2$ observables is shown in Table~\ref{tab:DF2}, where we quote the bounds for two representative sets of dimension-six operators: the left-left operators (present also in the SM) and operators with a different chirality, which arise in specific SM extensions. The bounds on the latter are stronger, especially in the kaon case, because of the larger hadronic matrix elements. The constraints related to CPV correspond to maximal phases, and are subject to the requirement that the NP contributions are smaller than $30\%$ ($60\%$) of the total contributions~\cite{NMFV} in the $B_d$ ($K$) system. Since the experimental status of CP violation in the $B_s$ system is not yet settled we simply require that the new physics contributions are smaller than the observed value of $\Delta m_{B_s}$ (for less naive treatments see {\it e.g.}~\cite{Bona:2007vi,ckmfitter}). \subsection{Correlations between $K$ and $D$ mixing} There are two different features that can provide flavor-related suppression factors: degeneracy and alignment. In general, low energy measurements can only constrain the product of these two suppression factors. An interesting exception occurs, however, for the LL operators of the type (\ref{eq:qlql}) where there is an independent constraint on the level of degeneracy \cite{Blum:2009sk}. We here briefly explain this point. Consider operators of the form \begin{equation}\label{xqoperator} \frac{1}{\Lambda_{\rm NP}^2}(\overline{Q}_{Li}(X_Q)_{ij}\gamma_\mu Q_{Lj}) (\overline{Q}_{Li}(X_Q)_{ij}\gamma^\mu Q_{Lj}), \end{equation} where $X_Q$ is an hermitian matrix. Without loss of generality, we can choose to work in the basis defined in Eq. (\ref{speint}): \begin{equation} Y^d=\lambda_d,\ \ \ Y^u=V^\dagger\lambda_u,\ \ \ X_Q=V_d^\dagger\lambda_Q V_d, \end{equation} where $\lambda_{Q}$ is a diagonal real matrix, and $V_d$ is a unitary matrix which parametrizes the misalignment of the operator (\ref{xqoperator}) with the down mass basis. The experimental constraints that are most relevant to our study come from $K^0$--$\overline{K^0}$ and $D^0$--$\overline{D^0}$ mixing, which involve only the first two generation quarks. When studying new physics effects, ignoring the third generation is often a good approximation to the physics at hand. Indeed, even when the third generation does play a role, our two generation analysis is applicable as long as there are no strong cancellations with contributions related to the third generation. In a two generation framework, $V$ depends on a single mixing angle (the Cabibbo angle $\theta_c$), while $V_d$ depends on a single angle and a single phase. To understand various aspects of our analysis, it is useful, however, to provisionally set the phase to zero, and study only CP conserving (CPC) observables. We thus have \begin{eqnarray}\label{twogen} \lambda_Q&=&{\rm diag}(\lambda_1,\lambda_2),\ \ \ V=\left(\begin{array}{cc} \cos\theta_c & \sin\theta_c \\ -\sin\theta_c & \cos\theta_c \end{array}\right),\ \ \ V_d=\left(\begin{array}{cc} \cos\theta_d & \sin\theta_d \\ -\sin\theta_d & \cos\theta_d \end{array}\right). \end{eqnarray} It is convenient to define \begin{equation} \lambda_{12}=\frac12(\lambda_{1}+\lambda_{2}),\ \ \delta_{12}=\frac{\lambda_{1}-\lambda_{2}}{\lambda_{1}+\lambda_{2}},\ \ \Lambda_{12}=\delta_{12}\lambda_{12}. \end{equation} Thus $\lambda_{12}$ parametrizes the overall, flavor-diagonal suppression of $X_Q$ (in particular, loop factors), $\delta_{12}$ parametrizes suppression that is coming from approximate degeneracy between the eigenvalues of $X_Q$, and $\theta_d$ and $\theta_c-\theta_d$ parametrize the suppression that comes from alignment with, respectively, the down and the up sector. The main point is the following: Alignment can entirely suppress the contribution to either $K^0$--$\overline{K^0}$ mixing ($\theta_d=0$) or $D^0$--$\overline{D^0}$ mixing ($\theta_d=\theta_c$) but not to both. Thus, the flavor measurements give a constraint on $\Lambda_{12}$ which reads \cite{Blum:2009sk} \begin{equation}\label{boucpc} \Lambda_{12}\leq3.8\times10^{-3} \left(\frac{\Lambda_{\rm NP}}{1\ {\rm TeV}}\right). \end{equation} If we switch on the CP violating phase $\gamma$ in $V_d$ then, for $0.03\mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$}|\sin\gamma|\lsim0.98$, we find \cite{Blum:2009sk} \begin{equation}\label{boucpv} \Lambda_{12}\leq\frac{4.8\times10^{-4}}{\sqrt{\sin2\gamma}} \left(\frac{\Lambda_{\rm NP}}{1\ {\rm TeV}}\right). \end{equation} We learn that, with a loop suppression of order $\lambda_{12}\sim\alpha_2$, the degeneracy should be stronger than $0.02$. \subsection{Top Physics} While the present direct determination of top-quark decay rates is at the ${\cal O}(1\%)$ level of accuracy~\cite{TevT}, orders of magnitude improvement is expected at the LHC. With $100\,{\rm fb}^{-1}$ of data, the LHC will be sensitive (at 95\% CL) to branching ratios of ${\cal O}\left(10^{-5}\right)$ in the $t\to u^i\,Z/\gamma$ channels~\cite{Carvalho:2007yi}, where $u^i=u,c$. Within the SM, these channels have branching ratios in the $10^{-13}$ range, so any experimental observation would be a clear sign of physics beyond the SM. The LHC sensitivity to additional $\Delta t=1$ processes, such as $t\to c\, G,\phi$, is more limited (see {\it e.g.}~Ref.~\cite{tfcncelse} and references therein). The impact of the projected LHC bounds on top flavor violation can be described in a model independent manner. The leading contributions to the above processes are mediated via a set of dimension-six operators~\cite{tEFT}, which can be classified according to the SM global flavor symmetries~\cite{Fox:2007in}. Focusing on the chirality of the quark fields, these can be written as ${\cal O}_{A^t B}$, where $A^t,B=L,R$ and $A^t$ denotes top-quark chirality. Bounds from $B$ physics can give severe constraints on the operators with $B=L$~\cite{Fox:2007in}. The situation changes, however, in case of a flavor alignment of these effective operators, such that they are diagonal in the down-type mass basis. In this limit no bounds can be derived on the ${\cal O}_{LR}$ and ${\cal O}_{RL}$ operators~\cite{GMP}. On the other hand, the ${\cal O}_{L L}$ operators cannot be aligned simultaneously with the down- and up-quark mass bases. As a result, combining bounds on $t\to u^i$ with bounds from $b\to s l^+ l^-$ we can get stringent constraints on the operator ${\cal O}^h_{LL}=\bar Q_i \gamma^\mu \left(X_Q\right)_{ij}Q_j\left(\phi^\dagger \overleftrightarrow{D}_\mu \phi\right)/\Lambda^2_t$. Normalising the $SU(3)_Q$ adjoint spurion, $X_Q= a_i T_i$, such that the projection onto the $SU(3)$ generators has a unit norm ($\sum_i a_i^2 =1$), we can indentify a unique combination which minimizes the combined bound from top and $B$ constraints. Taking into account the projected LHC sensitivity on $t\to u^i$ decays, we can expect to reach bounds of $\mathcal{O}(600\rm~GeV)$ on the scale $\Lambda_t$~\cite{GMP}, in the absence of a NP signal. \section{Minimal Flavor Violation} \label{sec:mfv} A very reasonable, although quite pessimistic, set up which avoids the new physics flavor problem is the so-called Minimal Flavor Violation (MFV) hypothesis. Under this assumption, flavor-violating interactions are linked to the known structure of Yukawa couplings also beyond the SM. As a result, non-standard contributions in FCNC transitions turn out to be suppressed to a level consistent with experiments even for $\Lambda \sim$~few TeV. One of the most interesting aspects of the MFV hypothesis is that it can naturally be implemented within the generic effective Lagrangian in Eq.~(\ref{eq:effL}). Furthermore, SM extensions where the flavor hierarchy is generated at a scale much higher than other dynamical scales tend to flow to the MFV class of models in the infra-red. The MFV hypothesis consists of two ingredients~\cite{D'Ambrosio:2002ex}: (i)~a {\em flavor symmetry} and (ii)~a set of {\em symmetry-breaking terms}. The symmetry is nothing but the large global symmetry of the SM Lagrangian in absence of Yukawa couplings shown in Eq.~(\ref{susuu}). Since this global symmetry, and particularly the ${\rm SU}(3)$ subgroups controlling quark flavor-changing transitions, is already broken within the SM, we cannot promote it to be an exact symmetry of the NP model. Some breaking would appear at the quantum level because of the SM Yukawa interactions. The most restrictive assumption we can make to {\em protect} in a consistent way quark-flavor mixing beyond the SM is to assume that $Y^d$ and $Y^u$ are the only sources of flavor symmetry breaking also in the NP model. To implement and interpret this hypothesis in a consistent way, we can assume that $SU(3)^3_{q}$ is a good symmetry and promote $Y^{u,d}$ to be non-dynamical fields (spurions) with non-trivial transformation properties under $SU(3)^3_{q}$: \begin{equation} Y^u \sim (3, \bar 3, 1)~,\qquad Y^d \sim (3, 1, \bar 3)~.\qquad \end{equation} If the breaking of the symmetry occurs at very high energy scales, at low-energies we would only be sensitive to the background values of the $Y$, {\it i.e.}~to the ordinary SM Yukawa couplings. Employing the effective-theory language, an effective theory satisfies the criterion of Minimal Flavor Violation in the quark sector\footnote{~The notion of MFV can be extended also to the letpon sector. However, in this case there is not a unique way to define the minimal sources of flavour symmetry breaking if we want to keep track of non-vanishing neutrino masses~\cite{Cirigliano:2005ck,Davidson:2006bd}.} if all higher-dimensional operators, constructed from SM and $Y$ fields, are invariant (formally) under the flavor group $SU(3)^3_{q}$~\cite{D'Ambrosio:2002ex}. Invariance under CP may or may not be imposed in addition. According to this criterion one should in principle consider operators with arbitrary powers of the (dimensionless) Yukawa fields. However, a strong simplification arises by the observation that all the eigenvalues of the Yukawa matrices are small, but for the top one (and possibly the bottom one, see later), and that the off-diagonal elements of the CKM matrix are very suppressed. Working in the basis in Eq.~(\ref{speint}), and neglecting the ratio of light quark masses over the top mass, we have \begin{equation} \left[ Y^u (Y^u)^\dagger \right]^n_{i\not = j} ~\approx~ y_t^{2n} V^*_{ti} V_{tj}~. \label{eq:basicspurion} \end{equation} As a consequence, including high powers of the the Yukawa matrices amounts only to a redefinition of the overall factor in (\ref{eq:basicspurion}) and the the leading $\Delta F=2$ and $\Delta F=1$ FCNC amplitudes get exactly the same CKM suppression as in the SM: \begin{eqnarray} \mathcal{A}(d^i \to d^j)_{\rm MFV} &=& (V^*_{ti} V_{tj})^{\phantom{a}} \mathcal{A}^{(\Delta F=1)}_{\rm SM} \left[ 1 + a_1 \frac{ 16 \pi^2 M^2_W }{ \Lambda^2 } \right]~, \\ \mathcal{A}(M_{ij}-{\overline{M}_{ij}})_{\rm MFV} &=& (V^*_{ti} V_{tj})^2 \mathcal{A}^{(\Delta F=2)}_{\rm SM} \left[ 1 + a_2 \frac{ 16 \pi^2 M^2_W }{ \Lambda^2 } \right]~, \label{eq:FC} \end{eqnarray} where the $\mathcal{A}^{(i)}_{\rm SM}$ are the SM loop amplitudes and the $a_i$ are $\mathcal{O}(1)$ real parameters. The $a_i$ depend on the specific operator considered but are flavor independent. This implies the same relative correction in $s\to d$, $b\to d$, and $b\to s$ transitions of the same type. Within the MFV framework, several of the constraints used to determine the CKM matrix (and in particular the unitarity triangle) are not affected by NP~\cite{Buras:2000dm}. In this framework, NP effects are negligible not only in tree-level processes but also in a few clean observables sensitive to loop effects, such as the time-dependent CPV asymmetry in $B_d \to \psi K_{L,S}$. Indeed the structure of the basic flavor-changing coupling in Eq.~(\ref{eq:FC}) implies that the weak CPV phase of $B_d$--$\overline{B}_d$ mixing is arg[$(V_{td}V_{tb}^*)^2$], exactly as in the SM. This construction provides a natural (a posteriori) justification of why no NP effects have been observed in the quark sector: by construction, most of the clean observables measured at $B$ factories are insensitive to NP effects in the MFV framework. \begin{table}[t] \begin{minipage}{\textwidth} \begin{center} \begin{tabular}{l|c|l} Operator & ~Bound on $\Lambda$~ & ~Observables \\ \hline\hline $H^\dagger \left( \overline{D}_R Y^{d\dagger} Y^u Y^{u\dagger} \sigma_{\mu\nu} Q_L \right) (e F_{\mu\nu})$ & ~$6.1$~TeV & ~$B\to X_s \gamma$, $B\to X_s \ell^+ \ell^-$\\ $\frac{1}{2} (\overline{Q}_L Y^u Y^{u\dagger} \gamma_{\mu} Q_L)^2 \phantom{X^{X^X}_{iii}}$ & ~$5.9$~TeV & ~$\epsilon_K$, $\Delta m_{B_d}$, $\Delta m_{B_s}$ \\ $H_D^\dagger \left( \overline{D}_R Y^{d\dagger} Y^u Y^{u\dagger} \sigma_{\mu\nu} T^a Q_L \right) (g_s G^a_{\mu\nu})$ &~$3.4$~TeV & ~$B\to X_s \gamma$, $B\to X_s \ell^+ \ell^-$\\ $\left( \overline{Q}_L Y^u Y^{u\dagger} \gamma_\mu Q_L \right) (\overline{E}_R \gamma_\mu E_R)$ & ~$2.7$~TeV & ~$B\to X_s \ell^+ \ell^-$, $B_s\to\mu^+\mu^-$ \\ $~i \left( \overline{Q}_L Y^u Y^{u\dagger} \gamma_\mu Q_L \right) H_U^\dagger D_\mu H_U$ &~$2.3$~TeV &~$B\to X_s \ell^+ \ell^-$, $B_s\to\mu^+\mu^-$\\ $\left( \overline{Q}_L Y^u Y^{u\dagger} \gamma_\mu Q_L \right) ( \overline{L}_L \gamma_\mu L_L)$ &~$1.7$~TeV & ~$B\to X_s \ell^+ \ell^-$, $B_s\to\mu^+\mu^-$\\ $\left( \overline{Q}_L Y^u Y^{u\dagger} \gamma_\mu Q_L \right) (e D_\mu F_{\mu\nu})$ &~$1.5$~TeV & ~$B\to X_s \ell^+ \ell^-$\\ \end{tabular} \end{center} \end{minipage} \caption{\label{tab:MFV} Bounds on the scale of new physics (at 95\% C.L.) for some representative $\Delta F=1$~\cite{Hurth:2008jc} and $\Delta F=2$~\cite{Bona:2007vi} MFV operators (assuming effective coupling $\pm 1/\Lambda^2$), and corresponding observables used to set the bounds.} \end{table} In Table~\ref{tab:MFV} we report a few representative examples of the bounds on the higher-dimensional operators in the MFV framework. For simplicity, only leading spurion dependence is shown on the left-handed column. The built-in CKM suppression leads to bounds on the effective scale of new physics not far from the TeV region. These bounds are very similar to the bounds on flavor-conserving operators derived by precision electroweak tests. This observation reinforces the conclusion that a deeper study of rare decays is definitely needed in order to clarify the flavor problem: the experimental precision on the clean FCNC observables required to obtain bounds more stringent than those derived from precision electroweak tests (and possibly discover new physics) is typically in the $1\%-10\%$ range. Although MFV seems to be a natural solution to the flavor problem, it should be stressed that (i) this is not a theory of flavor (there is no explanation for the observed hierarchical structure of the Yukawas), and (ii) we are still far from having proved the validity of this hypothesis from data (in the effective theory language we can say that there is still room for sizable new sources of flavor symmetry breaking beside the SM Yukawa couplings~\cite{Feldmann:2006jk}). A proof of the MFV hypothesis can be achieved only with a positive evidence of physics beyond the SM exhibiting the flavor-universality pattern (same relative correction in $s\to d$, $b\to d$, and $b\to s$ transitions of the same type) predicted by the MFV assumption. While this goal is quite difficult to be achieved, the MFV framework is quite predictive and thus could easily be falsified: in Table~\ref{tab:MFVbounds} we list some clean MFV predictions which could be falsified by future experiments. Violations of these bounds would not only imply physics beyond the SM, but also a clear signal of new sources of flavor symmetry breaking beyond the Yukawa couplings. \begin{table}[t] \begin{minipage}{\textwidth} \begin{center} \begin{tabular}{l|l|l|l} Observable & ~Experiment & ~{MFV prediction}~ & ~{SM prediction} \\ \hline \hline $ \beta_s$~from~$\mathcal{A}_{\rm CP}(B_s \to \psi \phi)$ & ~[0.10, 1.44] @ 95\% CL~ & ~$0.04(5)^*$ & ~$0.04(2)$ \\ \hline $\mathcal{A}_{\rm CP}(B \to X_s \gamma)$ & ~$< 6\%$ @ 95\% CL~ & ~$<0.02^*$ & ~$<0.01$ \\ \hline $\mathcal{B}(B_d \to \mu^+ \mu^-)$ & ~$<1.8 \times 10^{-8}$ \qquad & ~$<1.2 \times 10^{-9}$ & ~$1.3(3)\times 10^{-10}$ \\ \hline $\mathcal{B}(B \to X_s \tau^+ \tau^-)$ & ~ -- ~ & ~$< 5 \times 10^{-7}$ & ~$1.6(5)\times 10^{-7}$ \\ \hline $\mathcal{B}(K_L \to \pi^0 \nu \bar \nu)$ & ~$<2.6 \times 10^{-8}$ @ 90\% CL & ~$<2.9\times 10^{-10}$ & ~$2.9(5)\times 10^{-11}$ \\ \end{tabular} \end{center} \end{minipage} \caption{Some predictions derived in the MFV framework. Stars implies that the prediction is not valid in the GMFV case at large $\tan\beta$ (see Section~\ref{GMFV}). \label{tab:MFVbounds} } \end{table} The idea that the CKM matrix rules the strength of FCNC transitions also beyond the SM has become a very popular concept in recent literature and has been implemented and discussed in several works. It is worth stressing that the CKM matrix represents only one part of the problem: a key role in determining the structure of FCNCs is also played by quark masses, or by the Yukawa eigenvalues. In this respect, the MFV criterion provides the maximal protection of FCNCs (or the minimal violation of flavor symmetry), since the full structure of Yukawa matrices is preserved. At the same time, this criterion is based on a renormalization-group-invariant symmetry argument, which can be implemented independently of any specific hypothesis about the dynamics of the new-physics framework. This model-independent structure does not hold in most of the alternative definitions of MFV models that can be found in the literature. For instance, the definition of Ref.~\cite{Buras:2003jf} (denoted constrained MFV, or CMFV) contains the additional requirement that the effective FCNC operators playing a significant role within the SM are the only relevant ones also beyond the SM. This condition is realized only in weakly coupled theories at the TeV scale with only one light Higgs doublet, such as the MSSM with small $\tan\beta$ where no large logs, or sizable anomalous dimension are present~\cite{Kagan:2009bn} (see also \cite{Feldmann:2008ja}). It does not hold in several other frameworks, such as Higgsless models, 5D MFV models, or the MSSM with large $\tan\beta$. In NP models where sizable anomalous dimensions are present, the expansion in powers of the Yukawa spurions cannot be truncated at the first non-trivial order. In this limit, denoted as General MFV (GMFV), higher-order terms in the third-generation Yukawa couplings need to be re-summed. As shown in Eq.~(\ref{eq:basicspurion}), if only the top-quark Yukawa coupling is of order one, this resummation has negligible consequences for flavour-violating processes. (To measure the effect requires that the accuracy on rare $K$ ($D$) decays would become strong enough to probe the contributions related to the charm (strange) quark Yukawa coupling~\cite{Kagan:2009bn}). However, significant differences from a linear expansion may arise if both top- and bottom-quark Yukawa couplings are order one as it happens, for instance, in the large $\tan\beta$ regime. \subsection{MFV at large $\tan\beta$.} \label{eq:largetanb} If the Yukawa Lagrangian contains more than a single Higgs field, we can still assume that the Yukawa couplings are the only irreducible breaking sources of $SU(3)^3_{q}$, but we can change their overall normalization. A particularly interesting scenario is the two-Higgs-doublet model where the two Higgs fields, $\phi_U$ and $\phi_D$, are coupled separately to up- and down-type quarks: \begin{equation}\label{eq:LY2} - \mathcal{L}^{\rm 2HDM}_{\rm Yukawa} = Y^d_{ij}~{\overline{Q}_{Li}}\phi_D D_{Rj} +Y^u_{ij}~{\overline{Q}_{Li}} \phi_U U_{Rj}+Y^e_{ij}~\overline{L}_{Li}\phi_D E_{Rj} +{\rm h.c.}. \end{equation} This Lagrangian is invariant under an extra ${\rm U}(1)$ symmetry with respect to the one-Higgs Lagrangian in Eq.~(\ref{Hqint}): a symmetry under which the only charged fields are $D_R$ and $E_R$ (charge $+1$) and $\phi_D$ (charge $-1$). This symmetry, denoted ${\rm U}(1)_{\rm PQ}$, prevents tree-level FCNCs and implies that $Y^{u,d}$ are the only sources of $SU(3)^3_{q}$ breaking appearing in the Yukawa interaction (similar to the one-Higgs-doublet scenario). Consistently with the MFV hypothesis, we can then assume that $Y^{u,d}$ are the only relevant sources of $SU(3)_{q}^3$ breaking appearing in all the low-energy effective operators. This is sufficient to ensure that flavor-mixing is still governed by the CKM matrix, and naturally guarantees a good agreement with present data in the $\Delta F =2$ sector. However, the extra symmetry of the Yukawa interaction allows us to change the overall normalization of $Y^{u,d}$ with interesting phenomenological consequences in specific rare modes. These effects are related only to the large value of bottom Yukawa, and indeed can be found also in other NP frameworks where there is no extended Higgs sector, but the bottom Yukawa coupling is of order one~\cite{Kagan:2009bn}. Assuming the Lagrangian in Eq.~(\ref{eq:LY2}), the normalization of the Yukawa couplings is controlled by the ratio of the vacuum expectation values of the two Higgs fields, or by the parameter \begin{equation} \tan\beta = \langle \phi_U\rangle/\langle \phi_D\rangle~. \end{equation} For $\tan\beta\gg1 $ the smallness of the $b$ quark and $\tau$ lepton masses can be attributed to the smallness of $1/\tan\beta$ rather than to the corresponding Yukawa couplings. As a result, for $\tan\beta\gg1$ we cannot anymore neglect the down-type Yukawa coupling. Moreover, the ${\rm U}(1)_{\rm PQ}$ symmetry cannot be exact: it has to be broken at least in the scalar potential in order to avoid the presence of a massless pseudoscalar Higgs. Even if the breaking of ${\rm U}(1)_{\rm PQ}$ and $SU(3)^3_{q}$ are decoupled, the presence of ${\rm U}(1)_{\rm PQ}$ breaking sources can have important implications on the structure of the Yukawa interaction, especially if $\tan\beta$ is large~\cite{Hall:1993gn,Blazek:1995nv, Isidori:2001fv,D'Ambrosio:2002ex}. We can indeed consider new dimension-four operators such as \begin{equation} \epsilon~ \overline{Q}_L Y^d D_R \tilde \phi_U \qquad {\rm or} \qquad \epsilon~ \overline{Q}_L Y^uY^{u\dagger} Y^d D_R \tilde \phi_U~, \label{eq:O_PCU} \end{equation} where $\epsilon$ denotes a generic MFV-invariant ${\rm U}(1)_{\rm PQ}$-breaking source. Even if $\epsilon \ll 1 $, the product $\epsilon \times \tan\beta$ can be $\mathcal{O}(1)$, inducing large corrections to the down-type Yukawa sector: \begin{equation} \epsilon~ \overline{Q}_L Y^d D_R \tilde \phi_U \ \stackrel{vev}{\longrightarrow} \ \epsilon~ \overline{Q}_L Y^d D_R \langle \tilde \phi_U \rangle = (\epsilon\times\tan\beta)~ \overline{Q}_L Y^d D_R \langle \phi_D \rangle~. \end{equation} Since the $b$-quark Yukawa coupling becomes $\mathcal{O}(1)$, the large-$\tan\beta$ regime is particularly interesting for helicity-suppressed observables in $B$ physics. One of the clearest phenomenological consequences is a suppression (typically in the $10-50\%$ range) of the $B \to \ell \nu$ decay rate with respect to its SM expectation~\cite{Hou:1992sy}. Potentially measurable effects in the $10-30\%$ range are expected also in $B\to X_s \gamma$~\cite{Carena:1999py} and $\Delta M_{B_s}$~\cite{Buras:2001mb}. Given the present measurements of $B \to \ell \nu$, $B\to X_s \gamma$, and $\Delta M_{B_s}$, none of these effects seems to be favored by data. However, present errors are still sizable compared to the estimated NP effects. The most striking signature could arise from the rare decays $B_{s,d}\to \ell^+\ell^-$, whose rates could be enhanced over the SM expectations by more than one order of magnitude~\cite{Hamzaoui:1998nu}. An enhancement of both $B_{s}\to \ell^+\ell^-$ and $B_{d}\to \ell^+\ell^-$ respecting the MFV relation $\Gamma(B_{s}\to \ell^+\ell^-)/\Gamma(B_{d}\to \ell^+\ell^-) \approx |V_{ts}/V_{td}|^2$ would be an unambiguous signature of MFV at large $\tan\beta$~\cite{Hurth:2008jc}. \subsection{MFV with additional flavor-diagonal phases} \label{GMFV} The breaking of the $SU(3)_q^3$ flavor group and the breaking of the discrete CP symmetry are not necessarily related, and we can add flavor-diagonal CPV phases to generic MFV models~\cite{Ellis:2007kb,Mercolli:2009ns,Colangelo:2008qp}. Because of the experimental constraints on electric dipole moments (EDMs), which are generally sensitive to such flavour-diagonal phases~\cite{Mercolli:2009ns}, in this more general case the bounds on the scale of new physics are substantially higher with respect to the ``minimal'' case, where the Yukawa couplings are assumed to be the only breaking sources of both symmetries~\cite{D'Ambrosio:2002ex} If $\tan\beta$ is large, the inclusion of flavor-diagonal phases has interesting effects also in flavour-changing processes. The main consequences, derived in a model independent manner, can be summarized as follows~\cite{Kagan:2009bn}: (i) extra CPV can only arise from flavor diagonal CPV sources in the UV theory; (ii) the extra CP phases in $B_s-\bar B_s$ mixing provide an upper bound on the amount of CPV in $B_d-\bar B_d$ mixing; (iii) if operators containing right-handed light quarks are sub-dominant then the extra CPV is equal in the two systems, and is negligible in transitions between the first two generations to the third one. Conversely, these operators can break the correlation between CPV in the $B_s$ and $B_d$ systems, and can induce significant new CPV in $\epsilon_K$. \section{Supersymmetry} \label{sec:susy} Supersymmetric models provide, in general, new sources of flavor violation, for both the quark and the lepton sectors. The main new sources are the supersymmetry breaking soft mass terms for squarks and sleptons, and the trilinear couplings of a Higgs field with a squark-antisquark, or slepton-antislepton pairs. Let us focus on the squark sector. The new sources of flavor violation are most commonly analyzed in the basis in which the corresponding (down or up) quark mass matrix and the neutral gaugino vertices are diagonal. In this basis, the squark masses are not necessarily flavor-diagonal, and have the form \begin{equation} \tilde q_{Mi}^*(M_{\tilde q}^2)^{MN}_{ij}\tilde q_{Nj}= (\tilde q_{Li}^*\ \tilde q_{Rk}^*)\left(\begin{array}{cc} (M^2_{\tilde q})_{Lij} & A^q_{il}v_q \cr A^q_{jk}v_q & (M^2_{\tilde q})_{Rkl} \cr \end{array}\right) \left(\begin{array}{c} \tilde q_{Lj} \cr \tilde q_{Rl} \cr \end{array}\right), \end{equation} where $M,N=L,R$ label chirality, and $i,j,k,l=1,2,3$ are generation indices. $(M^2_{\tilde q})_L$ and $(M^2_{\tilde q})_R$ are the supersymmetry-breaking squark masses-squared. The $A^q$ parameters enter in the trilinear scalar couplings $A^q_{ij}\phi_q\widetilde q_{Li}\widetilde q_{Rj}^*$, where $\phi_q$ $(q=u,d)$ is the $q$-type Higgs boson and $v_q=\langle\phi_q\rangle$. In this basis, flavor violation takes place through one or more squark mass insertion. Each mass insertion brings with it a factor of $(\delta_{ij}^q)_{MN}\equiv(M^2_{\tilde q})^{MN}_{ij}/\tilde m_q^2$, where $\tilde m^2_q$ is a representative $q$-squark mass scale. Physical processes therefore constrain \begin{equation} [(\delta^q_{ij})_{MN}]_{\rm eff}\sim{\rm max}[(\delta^q_{ij})_{MN}, (\delta^q_{ik})_{MP}(\delta^q_{kj})_{PN},\ldots,(i\leftrightarrow j)]. \end{equation} For example, \begin{equation} [(\delta^d_{12})_{LR}]_{\rm eff}\sim{\rm max}[A^d_{12}v_d/\tilde m_d^2, (M^2_{\tilde d})_{L1k}A^d_{k2}v_d/\tilde m_d^4,A^d_{1k}v_d(M^2_{\tilde d})_{Rk2}/\tilde m_d^4,\ldots,(1\leftrightarrow2)]. \end{equation} Note that the contributions with two or more insertions may be less suppressed than those with only one. In terms of mass basis parameters, the $(\delta^q_{ij})_{MM}$'s stand for a combination of mass splittings and mixing angles: \begin{equation} (\delta^q_{ij})_{MM}=\frac{1}{\tilde m_{q}^2}\sum_\alpha (K^q_M)_{i\alpha}(K^q_M)_{j\alpha}^*\Delta\tilde m^2_{q_\alpha}, \end{equation} where $K^q_M$ is the mixing matrix in the coupling of the gluino (and similarly for the bino and neutral wino) to $q_{Li}-\tilde q_{M\alpha}$; $\tilde m^2_q=\frac13\sum_{\alpha=1}^3\tilde m_{q_{M\alpha}}^2$ is the average squark mass-squared, and $\Delta\tilde m^2_{q_\alpha}=\tilde m^2_{q_\alpha}-\tilde m^2_q$. Things simplify considerably when the two following conditions are satisfied \cite{Hiller:2008sv}, which means that a two generation effective framework can be used (for simplicity, we omit here the chirality index): \begin{equation} |K_{ik}K_{jk}^*|\ll|K_{ij}K_{jj}^*|,\ \ \ |K_{ik}K_{jk}^*\Delta\tilde m^2_{q_jq_i}|\ll|K_{ij}K_{jj}^*\Delta\tilde m^2_{q_jq_i}|, \end{equation} where there is no summation over $i,j,k$ and where $\tilde m^2_{q_jq_i}=\tilde m^2_{q_j}-\tilde m^2_{q_i}$. Then, the contribution of the intermediate $\tilde q_k$ can be neglected and, furthermore, to a good approximation, $K_{ii}K_{ji}^*+K_{ij}K_{jj}^*=0$. For these cases, we obtain \begin{equation}\label{eq:delmass} (\delta^q_{ij})_{MM}=\frac{\Delta\tilde m^2_{q_jq_i}}{\tilde m_{q}^2} (K^q_M)_{ij}(K^q_M)_{jj}^*. \end{equation} It is further useful to use instead of $\tilde m_q$ the mass scale $\tilde m^q_{ij}=\frac12(\tilde m_{q_i}+\tilde m_{q_j})$ \cite{Raz:2002zx}. We also define \begin{equation} \langle\delta^q_{ij}\rangle=\sqrt{(\delta_{ij}^{q})_{LL} (\delta^{q}_{ij})_{RR}}. \end{equation} The new sources of flavor and CP violation contribute to FCNC processes via loop diagrams involving squarks and gluinos (or electroweak gauginos, or higgsinos). If the scale of the soft supersymmetry breaking is below TeV, and if the new flavor violation is of order one, and/or if the phases are of order one, then these contributions could be orders of magnitude above the experimental bounds. Imposing that the supersymmetric contributions do not exceed the phenomenological constraints leads to constraints of the form $(\delta^q_{ij})_{MM}\ll1$. Such constraints imply that either quasi-degeneracy ($\Delta\tilde m^2_{q_jq_i}\ll\tilde m^{q2}_{ij}$) or alignment ($|K^q_{ij}|\ll1$) or a combination of the two mechanisms is at work. Table \ref{tab:exp} presents the constraints obtained in Refs. \cite{Masiero:2005ua,Ciuchini:2007cw,Buchalla:2008jp,Gedalia:2009kh} as appear in \cite{Hiller:2008sv}. Wherever relevant, a phase suppression of order 0.3 in the mixing amplitude is allowed, namely we quote the stronger between the bounds on ${\cal R}e(\delta^q_{ij})$ and $3{\cal I}m(\delta^q_{ij})$. The dependence of these bounds on the average squark mass $m_{\tilde q}$, the ratio $x\equiv m_{\tilde g}^2/m_{\tilde q}^2$ as well as the effect of arbitrary strong CP violating phases can be found in \cite{Hiller:2008sv}. \begin{table}[t] \begin{center} \begin{tabular}{cc|cc} \hline\hline \rule{0pt}{1.2em}% $q$\ & $ij\ $\ & $(\delta^{q}_{ij})_{MM}$ & $\langle\delta^q_{ij}\rangle$ \cr \hline $d$ & $12$\ & $\ 0.03\ $ & $\ 0.002\ $ \cr $d$ & $13$\ & $\ 0.2\ $ & $\ 0.07\ $ \cr $d$ & $23$\ & $\ 0.6\ $ & $\ 0.2\ $ \cr $u$ & $12$\ & $\ 0.1\ $ & $\ 0.008\ $ \cr \hline\hline \end{tabular} \caption{The phenomenological upper bounds on $(\delta_{ij}^{q})_{MM}$ and on $\langle\delta^q_{ij}\rangle$, where $q=u,d$ and $M=L,R$. The constraints are given for $m_{\tilde q}=1$ TeV and $x\equiv m_{\tilde g}^2/m_{\tilde q}^2=1$. We assume that the phases could suppress the imaginary parts by a factor $\sim0.3$. The bound on $(\delta^{d}_{23})_{RR}$ is about 3 times weaker than that on $(\delta^{d}_{23})_{LL}$ (given in table). The constraints on $(\delta^{d}_{12,13})_{MM}$, $(\delta^{u}_{12})_{MM}$ and $(\delta^{d}_{23})_{MM}$ are based on, respectively, Refs. \cite{Masiero:2005ua}, \cite{Ciuchini:2007cw} and \cite{Buchalla:2008jp}. \label{tab:exp}} \end{center} \end{table} For large $\tan\beta$, some constraints are modified from those in Table~\ref{tab:exp}. For instance, the effects of neutral Higgs exchange in $B_s$ and $B_d$ mixing give, for $\tan \beta =30$ and $x=1$ (see \cite{Hiller:2008sv,Foster:2006ze} and references therein for details): \begin{equation} \label{eq:bmixbounds} \langle \delta^d_{13}\rangle < 0.01 \cdot \left( \frac{M_{A^0}}{200 \, \mbox{GeV}} \right) , ~~~~~ \langle \delta^d_{23} \rangle < 0.04 \cdot \left( \frac{M_{A^0}}{200 \, \mbox{GeV}} \right) , \end{equation} where $M_{A^0}$ denotes the pseudoscalar Higgs mass, and the above bounds scale roughly as $(30/\tan \beta)^2$. The experimental constraints on the $(\delta^q_{ij})_{LR}$ parameters in the quark-squark sector are presented in Table~\ref{tab:expLRme}. The bounds are the same for $(\delta^q_{ij})_{LR}$ and $(\delta^q_{ij})_{RL}$, except for $(\delta^d_{12})_{MN}$, where the bound for $MN=LR$ is 10 times weaker. Very strong constraints apply for the phase of $(\delta^q_{11})_{LR}$ from EDMs. For $x=4$ and a phase smaller than 0.1, the EDM constraints on $(\delta^{u,d,\ell}_{11})_{LR}$ are weakened by a factor $\sim6$. \begin{table}[t] \label{tab:expLRme} \begin{center} \begin{tabular}{cc|c} \hline\hline \rule{0pt}{1.2em}% $q$\ & $ij$\ & $(\delta^{q}_{ij})_{LR}$\cr \hline $d$ & $12$\ & $\ 2\times10^{-4} \ $ \cr $d$ & $13$\ & $\ 0.08 \ $ \cr $d$ & $23$\ & $\ 0.01 \ $ \cr $d$ & $11$\ & $4.7\times10^{-6}$ \cr $u$ & $11$\ & $9.3\times 10^{-6}$ \cr $u$ & $12$\ & $\ 0.02 \ $\cr \hline\hline \end{tabular} \caption{The phenomenological upper bounds on chirality-mixing $(\delta_{ij}^{q})_{LR}$, where $q=u,d$. The constraints are given for $m_{\tilde q}=1$ TeV and $x\equiv m_{\tilde g}^2/m_{\tilde q}^2=1$. The constraints on $\delta^{d}_{12,13}$, $\delta^{u}_{12}$, $\delta^{d}_{23}$ and $\delta^{q}_{ii}$ are based on, respectively, Refs.~\cite{Masiero:2005ua}, \cite{Ciuchini:2007cw}, \cite{Buchalla:2008jp} and \cite{Raidal:2008jk} (with the relation between the neutron and quark EDMs as in \cite{Gabbiani:1996hi}).} \end{center} \end{table} While, in general, the low energy flavor measurements constrain only the combinations of the suppression factors from degeneracy and from alignment, such as Eq. (\ref{eq:delmass}), an interesting exception occurs when combining the measurements of $K^0$--$\overline{K^0}$ and $D^0$--$\overline{D^0}$ mixing to test the first two generation squark doublets. Here, for masses below the TeV scale, some level of degeneracy is unavoidable \cite{Blum:2009sk}: \begin{equation} \frac{m_{\widetilde Q_2}-m_{\widetilde Q_1}}{m_{\widetilde Q_2}+m_{\widetilde Q_1}}\leq\begin{cases} 0.034 & {\rm maximal\ phases} \cr 0.27 & {\rm vanishing\ phases}\cr \end{cases} \end{equation} The strong constraints in Tables \ref{tab:exp} and \ref{tab:expLRme} can be satisfied if the mediation of supersymmetry breaking to the MSSM is MFV. In particular, if at the scale of mediation, the supersymmetry breaking squark masses are universal, and the A-terms vanish or are proportional to the Yukawa couplings, then the model is phenomenologically safe. Indeed, there are several known mechanisms of mediation that are MFV (see, {\it e.g.} \cite{Shadmi:1999jy}). In particular, gauge-mediation \cite{Dine:1993yw,Dine:1994vc,Dine:1995ag,Meade:2008wd}, anomaly-mediation \cite{Randall:1998uk,Giudice:1998xp}, and gaugino-mediation \cite{Chacko:1999mi} are such mechanisms. (The renormalization group flow in the MSSM with generic MFV soft-breaking terms at some high scale has recently been discussed in Ref.~\cite{Colangelo:2008qp,Paradisi:2008qh}.) On the other hand, we do not expect gravity-mediation to be MFV and it could provide subdominant, yet observable flavor and CP violating effects \cite{Feng:2007ke}. \section{Extra Dimensions} \label{sec:exdim} Models of extra dimensions come in a large variety and the corresponding phenomenology, including the implications for flavor physics, changes from one extra dimension framework to another. Yet, as in the supersymmetric case, one can classify the new sources of flavor violation which generically arise: {\bf Bulk masses} - If the SM fields propagate in the bulk of the extra dimensions they can have bulk, vector-like, masses. These mass terms are of particular importance to flavor physics since they induce fermion localization which may yield hierarchies in the low energy effective couplings. Furthermore, the bulk masses, which define the extra dimension interaction basis, do not need to commute with the Yukawa matrices, and hence might induce contributions to FCNC processes, similarly to the squark soft masses-squared in supersymmetry. {\bf Cutoff, UV physics} - Since, generically, higher dimensional field theories are non-renormalizable, they rely on unspecified microscopic dynamics to provide UV completion of the models. Hence, they can be viewed as effective field theories and the impact of the UV physics is expected to be captured by a set of operators suppressed by the, framework dependent, cutoff scale. Without precise knowledge of the short distance dynamics, the additional operators are expected to carry generic flavor structure and contribute to FCNC processes. This is somewhat similar to ``gravity mediated'' contributions to supersymmetry-breaking soft terms which are generically expected to have an anarchic flavor structure and are suppressed by the Planck scale. {\bf ``Brane'' localized terms} - The extra dimensions have to be compact and typically contain defects and boundaries of smaller dimensions [in order, for example, to yield a chiral low energy four dimension (4D) theory]. These special points might contain different microscopical degrees of freedom. Therefore, generically, one expects that a different and independent class of higher dimension operators may be localized to this singular region in the extra dimension manifold. (These are commonly denoted `brane terms' even though, in most cases, they have very little to do with string theory). The brane-localized terms can, in principle, be of anarchic flavor structure and provide new flavor and CP violating sources. One important class of such operators are brane kinetic terms: their impact is somewhat similar to that of non-canonical kinetic terms which generically arise in supersymmetric flavor models. We focus on flavor physics of five dimension (5D) models, with bulk SM fields, since most of the literature focuses on this class. Furthermore, the new flavor structure that arises in 5D models captures most of the known effects of extra-dimension flavor models. Assuming a flat extra dimension, the energy range, $\Lambda_{\rm 5D} R$ (where $\Lambda_{\rm 5D}$ is the 5D effective cutoff scale and $R$ is the extra dimension radius with the extra dimension coordinate $y\in(0,\pi R)$), for which the 5D effective field theory holds, can be estimated as follows. Since gauge couplings in extra dimensional theories are dimensional, {\it i.e.}~$\alpha_{\rm 5D}$ has mass dimension $-1$, a rough guess (which is confirmed, up to order one corrections, by various NDA methods) is~\cite{Kribs:2006mq} $\Lambda_{\rm 5D} \sim 4 \pi/{\alpha_{\rm 5D}} \,.$ Matching this 5D gauge coupling to a 4D coupling of the SM, at leading order, ${1}/{g^2} = \pi R/{g_{\rm 5D}^2}\,,$ we obtain \begin{equation} \Lambda_{\rm 5D} R \sim \frac{4}{\alpha} \sim 30\,. \end{equation} Generically, the mass of the lightest Kaluza-Klein (KK) states, ${M_{\rm KK}}$, is of ${\cal O}\big(R^{-1}\big)$. If the extra dimension theory is linked to the solution of the hierarchy problem and/or directly accessible to near future experiments, then $R^{-1}= {\cal O}\big(\rm TeV\big)$. This implies an upper bound on the 5D cutoff: \begin{equation} \Lambda_{\rm 5D}\lesssim 10^2\,{\rm TeV}\ll \Lambda_K\sim 2 \times 10^5\,\rm TeV \,, \end{equation} where $ \Lambda_K$ is the scale required to suppress the generic contributions to $\epsilon_K$, discussed above (see Table~\ref{tab:DF2}). The above discussion ignores the possibility of splitting the fermions in the extra dimension. In split fermion models different bulk masses are assigned to different generations. Consequently fermions are localized and separated in the bulk of the extra dimension in a manner which may successfully address the SM flavor puzzle~\cite{ArkaniHamed:1999dc}. Separation in the extra dimension may suppress the contributions to $\epsilon_K$ from the higher-dimension cut-off induced operators. As shown in Table~\ref{tab:DF2}, the most dangerous operator is \begin{equation} {O^4_K} = \frac{1}{\Lambda_{\rm 5D}^2} \left(\bar s_L \,d_R\right) \left(\bar s_R\, d_L\right)~. \label{eq:O4K} \end{equation} This operator contains $s$ and $d$ fields of both chiralities. As a result, in a large class of split fermion models, the overlap suppression would be similar to that accounting for the smallness of the down and strange 4D Yukawa couplings. Integrating over the 5D profiles of the four quarks, this may yield a suppression factor of ${\cal O}\big(m_d m_s/v^2\big)\sim10^{-9}$. Together with the naive scale suppression, $1/\Lambda_{\rm 5D}^2$, the coefficient of ${O^4_K}$ can be sufficiently suppressed to be consistent with the experimental bound. In the absence of large brane kinetic terms (BKTs), however, fermion localization generates order one non-universal couplings to the gauge KK fields~\cite{Delgado:1999sv}. (See {\it e.g.}~\cite{AgasheTasi} and references therein. The case with large BKTs is similar to the warped case discussed below.) The fact that the bulk masses are, generically, not aligned with the 5D Yukawa couplings implies that KK gluon exchange processes induce, among others, the following operator in the low energy theory: $\left[(D_L)_{12}^2/(6{M^2_{\rm KK}})\right]\left(\bar s_L\, d_L\right)^2$, where $\left(D_{L}\right)_{12}\sim \lambda$ is the left-handed down-quark rotation matrix from the 5D interaction basis to the mass basis. This structure provides only a mild suppression to the resulting operator. It implies that to satisfy the $\epsilon_K$ constraint the KK and the inverse compactification scales have to be above $10^3$\,TeV, beyond the direct reach of near future experiments, and too high to be linked to a solution of the hierarchy problem. This problem can be solved by tuning the 5D flavor parameters and imposing appropriate 5D flavor symmetries to make the tuning stable. Once the 5D bulk masses are aligned with the 5D Yukawa matrices the KK gauge contributions would vanish and such a configuration is radiatively stable. The warped extra-dimension [Randall-Sundrum (RS)] framework~\cite{Randall:1999ee} provides a solution to the hierarchy problem. Moreover, with SM fermions propagating in the bulk, both the SM and the NP flavor puzzles can be addressed. The light fermions could be localized away from the TeV brane~\cite{GN}, where the Higgs is localized. Such a configuration can generate the observed Yukawa hierarchy and, at the same time, ensure that higher-dimensional operators are suppressed by a high cutoff scale associated with the location of the light fermions in the extra dimension~\cite{RSoriginal}. Furthermore, since the KK states are localized near the TeV brane, the couplings between the SM quarks and the gauge KK fields exhibit the hierarchical structure associated with SM masses and CKM mixings. This hierarchy in the couplings provides an extra protection against non-standard flavor-violating effects~\cite{Huber:2003tu} denoted as RS-GIM mechanism~\cite{aps} (see also~\cite{Burdman:2002gr}). It is interesting to note that an analogous mechanism is at work in models with strong dynamics at the TeV scale, with large anomalous dimension and partial compositeness~\cite{GK}. The link with strongly-interacting models in indeed motivated by the AdS/CFT correspondence~\cite{AdSCFT}, which implies that the above 5D framework is the dual description of 4D composite Higgs models~\cite{AdSCFTpheno}. Concerning the quark zero modes, the flavor structure of the above models as well as the phenomenology can be captured by using the following simple rules~\cite{aps,Gedalia:2009ws,Contino:2006nn}. In the 5D interaction basis, where the bulk masses $k\, C^{ij}_{x}$ are diagonal ($x=Q,U,D$; $i,j=1,2,3$; $k$ is the AdS curvature), the value $f_{x^i}$ of the profile of the quark zero modes is given by \begin{equation} \label{fs} f_{x^i}^2=(1-2c_{x^i})/( 1-\epsilon^{ 1-2c_{x^i} })\,. \end{equation} Here $c_{x^i}$ are the eigenvalues of the $C_x$ matrices, $\epsilon=\exp[-\xi]$, $\xi=\log[M_{\rm \overline{Pl}}/{\rm TeV}]$, and $M_{\rm \overline{Pl}}$ is the reduced Planck mass. If $c_{x^i}<1/2$, then $f_{x^i}$ is exponentially suppressed. Hence, order one variations in the 5D masses yield large hierarchies in the 4D flavor parameters. We consider the cases where the Higgs VEV either propagates in the bulk~\cite{Agashe:2008uz} or is localized on the IR brane. For a bulk Higgs case, the profile is given by $\tilde{v}( \beta, z) \simeq v\sqrt{k(1+\beta)} \bar z^{2+\beta}/ \epsilon $, where $\bar z\in(\epsilon,1)$ ($\bar z=1$ on the IR brane), and $\beta\geq0$. The $\beta=0$ case describes a Higgs maximally-spread into the bulk (saturating the AdS stability bound~\cite{Breitenlohner:1982bm}). The relevant part of the effective 4D Lagrangian, which involves the zero modes and the first KK gauge states can be approximated by~\cite{aps,Gedalia:2009ws} \begin{equation} \label{lagrangian}\hspace*{-.15cm} \mathcal{L}^{4D} \supset (Y^{u,d}_{\rm 5D})_{ij} \phi^{u,d} \, \bar Q_i f_{Q_i} \left(U,D\right)_j f_{U_j,D_j} r^\phi_{00} (\beta,c_{Q_i},c_{U_j,D_j}) +g_* G^1 x^\dagger_i x_i \left[f_{x^i}^2 {r^g_{00}}(c_{x^i}) -{1}/{\xi} \right] , \end{equation} where $\phi^{u,d}=\tilde \phi, \phi$, $g_* $ stands for a generic effective gauge coupling and summation over $i,j$ is implied. The correction for the couplings from the case of fully IR-localized KK and Higgs states is given by the functions $r^\phi_{00}$~\cite{Gedalia:2009ws} and ${r^g_{00}}$~\cite{cfw1,Csaki:2009bb}: \begin{equation} \label{r00} r^\phi_{00}(\beta,c_L,c_R) \approx \frac{\sqrt{2(1+\beta)}}{2+\beta-c_L-c_R} \,, \ \ \ {r^g_{00}}(c) \approx \frac{\sqrt{2}}{J_1(x_1)} \frac{0.7}{6-4c} \left( 1+e^{c/2} \right) \,, \end{equation} where $r^\phi_{00}(\beta,c_L,c_R)=1$ for brane-localized Higgs and $x_1 \approx 2.4$ is the first root of the Bessel function, $J_0(x_1)=0$. In Table~\ref{fstab} we present an example of a set of $f_{x^i}$-values that, starting from anarchical 5D Yukawa couplings, reproduce the correct hierarchy of the flavor parameters. We assume, for simplicity, an IR localized Higgs. The values depend on two input parameters: $f_{U^3}$, which has been determined assuming a maximally localized $t_R$ ($c_{u^3}=-0.5$), and ${y_{\rm 5D}}$, the overall scale of the 5D Yukawa couplings in units of $k$, which has been fixed to its maximal value assuming three KK states. On general grounds, the value of ${y_{\rm 5D}}$ is bounded from above, as a function of the number of KK levels, by the requirement that Yukawa interactions are perturbative below the cutoff of the theory, $\Lambda_{\rm 5D}\sim {N_{\rm KK}} k$, and it is bounded from below in order to account for the large top mass. Hence the following range for ${y_{\rm 5D}}$ is obtained (see {\it e.g.}~\cite{Agashe:2008uz,LCP}): \begin{equation} \frac{1}{ 2}\lesssim {y_{\rm 5D}} \lesssim \frac{2\pi}{ {N_{\rm KK}}} {\rm \ \ for \ brane \ Higgs\,;} \ \ \ \ \ \frac{1}{ 2}\lesssim {y_{\rm 5D}} \lesssim \frac{4\pi}{ \sqrt{N_{\rm KK}}} {\rm \ \ for \ bulk \ Higgs\,,} \label{lam5D}\end{equation} where we use the rescaling $y_{\rm 5D}\to y_{\rm 5D}\, \sqrt{1+\beta}$, which produces the correct $\beta\to \infty$ limit~\cite{HFCNC} and avoids subtleties in the $\beta=0$ case. {\small \begin{table}[t]\begin{center} \begin{tabular}{c|c|c|c} \hline\hline { Flavor}& { $f_Q$} & { $f_U$} & { $f_D$}\cr \hline 1 &$ {{A \lambda^{3}} { f_{Q^3}}}\sim 3 \times 10^{-3}$& $\frac{m_u}{m_t}\, \frac{ f_{U^3} }{ A \lambda^3} \sim1\times10^{-3}$& $\frac{m_d}{m_b}\, \frac{ f_{D^3} }{ A \lambda^3} \sim 2\times 10^{-3}$ \cr 2&$ {{ A \lambda^{2}} { f_{Q^3}}}\sim 1\times 10^{-2}$& $\frac{ m_c }{ m_t} \, \frac{ f_{U^3} }{ A \lambda^2} \sim 0.1$& $\frac{ m_s}{ m_b}\, \frac{ f_{D^3} }{ A \lambda^2} \sim 1\times 10^{-2}$ \cr 3 &$ \frac{ m_t}{ v {y_{\rm 5D}} f_{U^3}}\sim 0.3$ &$\sqrt2$& $\frac{ m_b }{ m_t} \, { f_{U^3}}\sim 2\times 10^{-2}$ \vspace*{.05cm}\cr \hline\hline\end{tabular} \caption{{\small Values of the $f_{x^i}$ parameters [Eq.~(\ref{fs})] which reproduce the observed quark masses and CKM mixing angles starting from anarchical 5D Yukawa couplings. We fix $f_{U^3}=\sqrt2$ and ${y_{\rm 5D}}=2$ (see text).}}\label{fstab} \end{center} \end{table} } \begin{table}[t] \begin{center} \begin{tabular}{c|c c|c c} \hline\hline \rule{0pt}{1.2em}% Observable & \multicolumn{2}{c}{$M_G^{\rm min}[\mathrm{TeV}]$} & \multicolumn{2}{|c}{${y_{\rm 5D}^{\rm min}}\ {\rm or}\ f_{Q_3}^{\rm max}$} \cr & IR Higgs& $\beta=0$ & IR Higgs & $\beta=0$ \cr \hline CPV-$B_d^{LLLL}$ & $12 f_{Q_3}^2$ &$12 f_{Q_3}^2$& $f_{Q_3}^{\rm max}=0.5$ &$f_{Q_3}^{\rm max}=0.5$\cr CPV-$B_d^{LLRR}$ & ${4.2/ {y_{\rm 5D}}}$ &${2.4/ {y_{\rm 5D}}}$& ${y_{\rm 5D}^{\rm min}}=1.4$ &${y_{\rm 5D}^{\rm min}}=0.82$ \cr CPV-$D^{LLLL}$ & $0.73 f_{Q_3}^2$ &$0.73 f_{Q_3}^2$& no bound & no bound\cr CPV-$D^{LLRR}$ & ${4.9/ {y_{\rm 5D}}}$ &${2.4/ {y_{\rm 5D}}}$& ${y_{\rm 5D}^{\rm min}}=1.6$ &${y_{\rm 5D}^{\rm min}}=0.8$ \cr $\epsilon_K^{LLLL}$ & $7.9 f_{Q_3}^2$& $7.9 f_{Q_3}^2$ & $f_{Q_3}^{\rm max}=0.62$ &$f_{Q_3}^{\rm max}=0.62$\cr $\epsilon_K^{LLRR}$ & ${49/ {y_{\rm 5D}}}$ &$ {24/ {y_{\rm 5D}}}$& above (\ref{lam5D}) & ${y_{\rm 5D}^{\rm min}}=8 $ \cr \hline \hline \end{tabular} \caption{Most significant flavor constraints in the RS framework. The values of ${y_{\rm 5D}^{\rm min}}$ and $ f_{Q_3}^{\rm max}$ correspond to ${M_{\rm KK}}=3$ TeV. The bounds are obtained assuming maximal CPV phases and $g_{s*}=3$. Entries marked `above (\ref{lam5D})' imply that for ${M_{\rm KK}}=3$ TeV, ${y_{\rm 5D}}$ is outside the perturbative range. \label{tab:rszi}} \end{center} \end{table} With anarchical 5D Yukawa matrices, an RS residual little CP problem remains~\cite{LCP}: Too large contributions to the neutron electric dipole moment (EDM)~\cite{aps}, and sizable chirally enhanced contributions to $\epsilon_K$~\cite{cfw1,Blanke:2008zb,Davidson:2007si,Bona:2007vi,RSC} are predicted. The RS leading contribution to $\epsilon_K$ is generated by a tree-level KK-gluon exchange which leads to an effective coupling for the chirality-flipping operator in (\ref{eq:O4K}) of the type~\cite{cfw1,Blanke:2008zb,Davidson:2007si,RSC} \begin{eqnarray}\label{eq:C4K} C_4^K &\simeq&\frac{g_{s*}^2}{M_{KK}^2} f_{Q_2} f_{Q_1} f_{d_2} f_{d_1} {r^g_{00}}(c_{Q_2}) {r^g_{00}}(c_{d_2}) \nonumber\\ &\sim& \frac{g_{s*}^2}{M_{KK}^2} \frac{2 m_d m_s }{(v {y_{\rm 5D}})^2} \frac{{r^g_{00}}(c_{Q_2}) {r^g_{00}}(c_{d_2})}{r^H_{00}(\beta,c_{Q_1},c_{d_1}) r^H_{00}(\beta,c_{Q_2},c_{d_2})}\,. \end{eqnarray} The final expression is independent of the $f_{x^i}$, so the bound in Table~\ref{tab:DF2} can be translated into constraints in the ${y_{\rm 5D}}-{M_{\rm KK}}$ plane. The analogous effects in the $D$ and $B$ systems yield numerically weaker bounds. Another class of contributions, which involves only left-handed quarks, is also important to constrain the $f_Q-{M_{\rm KK}}$ parameter space. In Table~\ref{tab:rszi} we summarize the resulting constraints. For the purpose of a quantitative analysis we set $g_{s*}=3$, as obtained by matching to the 4D coupling at one-loop~\cite{Agashe:2008uz} (for the impact of a smaller RS volume see~\cite{Davoudiasl:2008hx}). The constraints related to CPV correspond to maximal phases, and are subject to the requirement that the RS contributions are smaller than $30\%$ ($60\%$) of the SM contributions~\cite{NMFV} in the $B_d$ ($K$) system. The analytical expressions in the table have roughly a 10\% accuracy over the relevant range of parameters. Contributions from scalar exchange, either Higgs~\cite{HFCNC} or radion~\cite{Azatov:2008vm}, are not included since these are more model dependent and known to be weaker~\cite{Duling:2009pj} in the brane localized Higgs case. Constraints from $\epsilon'/\epsilon_K$ have a different parameter dependence than the $\epsilon_K$ constraints. Explicitly, for $\beta=0$, the $\epsilon'/\epsilon_K$ constraint reads $M_G^{\rm min}=1.2{y_{\rm 5D}}$ TeV. When combined with the $\epsilon$ constraint, we find $M_G^{\rm min}=5.5$ TeV with a corresponding ${y_{\rm 5D}^{\rm min}}=4.5$~\cite{Gedalia:2009ws}. The constraints summarized in Table~\ref{tab:rszi} and the contributions to the neutron EDM which generically require ${M_{\rm KK}}> {\cal O}\left(20\,\rm TeV\right)$~\cite{aps} are a clear manifestation of the RS little CP problem. The problem can be amended by various alignment mechanisms~\cite{LCP,Csaki:2009bb,Align}. In this case the bounds from the up sector, especially from CPV in the $D$ system~\cite{Blum:2009sk,Gedalia:2009kh}, become important. Constraint from $\Delta F=1$ processes (in either the down sector~\cite{aps,RSC} or $t\to c Z$~\cite{Agashe:2006wa}) are not included here, since they are weaker and, furthermore, these contributions can be suppressed (see~\cite{RSC}) due to incorporation of a custodial symmetry~\cite{Agashe:2006at}. \section{Future Prospects} \label{sec:future} The new physics flavor puzzle is the question of why, and in what way, the flavor structure of TeV-scale new physics is non-generic. Indeed, the flavor predictions of most new physics models are not a consequence of their generic features but rather of the special structures that are imposed specifically to satisfy the existing severe flavor bounds. Therefore, flavor physics is a powerful indirect probe of new physics. We hope that new physics not far above the weak scale will be discovered at the LHC. A major issue will then be to understand its flavor structure. While it is not easy to directly probe this flavor structure at high energy, a lot can be learned from low energy flavor physics. The precision with which we can probe high scale physics in flavor physics experiments is often limited by theoretical uncertainties. Moreover, in case of theoretically clean observables, the sensitivity to the new-physics scale increases slowly with the statistics of the experiment. Thus, the important questions in view of future experiments are the following: \begin{enumerate} \item What are the expected deviations from the SM predictions induced by new physics at the TeV scale? \item Which observables are not limited by theoretical uncertainties? \item In which case we can expect a substantial improvement on the experimental side? \item What will the measurements teach us if deviations from the SM are [not] seen? \end{enumerate} These questions have been analysed in a series of recent works (see e.g.~\cite{Raidal:2008jk,Buchalla:2008jp,Grossman:2009dw,Antonelli:2009ws,Buras:2009if}) and the main conclusions can be summarized as follows: \begin{enumerate} \item The expected deviations from the SM predictions induced by new physics at the TeV scale with generic flavor structure are already ruled out by many orders of magnitudes. On general grounds, we can expect any size of deviation below the current bounds. In the most pessimistic frameworks, such as MFV, the typical size of the deviations is at the few \% level in FCNC amplitudes. \item The theoretical limitations are highly process dependent. In most multi-hadron final states the calculation of decay amplitudes are already limited by theoretical uncertainties. However, several channels involving leptons in the final state, and selected time-dependent asymmetries, have a theoretical errors well below the current experimental sensitivity. \item On the experimental side there are good prospect of improvements. As summarized in Table~\ref{tab:future}, one order of magnitude improvements in several clean $B_{s,d}$, $D$, and $K$ observables are possible within a few years. Moreover, improvements of several orders of magnitudes are expected in top decays, which will be explored for the first time in great detail at the LHC. \item There is no doubt that new low-energy flavor data will be complementary with the high-$p_T$ part of the LHC program. As illustrated in the previous sections, the synergy of both data sets can teach us a lot about the new physics at the TeV scale. \end{enumerate} {\small \begin{table}[pt] \begin{tabular}{l|c|c|c|c|l}\hline & SM & Theory & Present & Future & Future \\ [-8 pt] \raisebox{10pt}{Observable} & prediction & error & result & error & Facility \\ \hline\hline $|V_{us}|\quad$ [$K\to\pi\ell\nu$] & input & $0.5\% \to 0.1\%_{\rm Latt}$ & $0.2246\pm0.0012$ & 0.1\% & {\footnotesize $K$ factory} \\ $|V_{cb}|\quad$ [$B\to X_c \ell \nu$] & input & $1\%$ & $(41.54 \pm 0.73) \times 10^{-3}$ & $1\%$ & {\footnotesize Super-$B$} \\ $|V_{ub}|\quad$ [$B\to \pi\ell\nu$] & input & $10\% \to 5\%_{\rm Latt}$ & $(3.38 \pm 0.36)\times 10^{-3}$ & 4\% & {\footnotesize Super-$B$ } \\ $\gamma\qquad\ $ [$B \to DK$] & input & $<1^\circ$ & $(70^{+27}_{-30})^\circ$ & $3^\circ$ & {\footnotesize LHCb} \\ \hline $S_{B_d \to \psi K}$ & $\sin(2\beta)$ & $\mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 0.01$ & $0.671 \pm 0.023$ & 0.01 & {\footnotesize LHCb} \\ $S_{B_s\to \psi\phi}$ & 0.036 & $\mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 0.01$ & $0.81^{+0.12}_{-0.32}$ & $0.01$ & {\footnotesize LHCb} \\ $S_{B_d \to \phi K}$ & $\sin(2\beta)$ & $\mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 0.05$ & $0.44 \pm 0.18$ & 0.1 & {\footnotesize LHCb} \\ $S_{B_s\to \phi\phi}$ & $0.036$ & $\mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 0.05$ & --- & $0.05$ & {\footnotesize LHCb} \\ $S_{B_d \to K^* \gamma}$ & few $\times$ 0.01 & $0.01$ & $ -0.16 \pm 0.22$ & $0.03$ & {\footnotesize Super-$B$} \\ $S_{B_s\to\phi\gamma}$ & few $\times$ 0.01 & $0.01$ & --- & $0.05$ & {\footnotesize LHCb} \\ $A_{\rm SL}^d$ & $-5 \times 10^{-4}$ & $10^{-4}$ & $-(5.8 \pm 3.4) \times 10^{-3}$ & $10^{-3}$ & {\footnotesize LHCb} \\ $A_{\rm SL}^s$ & $2 \times 10^{-5}$ & $< 10^{-5}$ & $(1.6 \pm 8.5) \times 10^{-3}$ & $10^{-3}$ & {\footnotesize LHCb} \\ \hline $A_{CP}(b\to s\gamma)$ & $<0.01$ & $<0.01$ & $-0.012\pm0.028$ & 0.005 & {\footnotesize Super-$B$} \\ $\mathcal{B}(B\to \tau\nu)$ & $1\times 10^{-4}$ & $20\% \to 5\%_{\rm Latt}$ & $(1.73\pm0.35)\times10^{-4}$ & 5\% & {\footnotesize Super-$B$} \\ $\mathcal{B}(B\to \mu\nu)$ & $4\times 10^{-7}$ & $20\% \to 5\%_{\rm Latt}$ & $<1.3\times10^{-6}$ & 6\% & {\footnotesize Super-$B$} \\ $\mathcal{B}(B_s\to \mu^+\mu^-)$ & $3\times 10^{-9}$ & $20\% \to 5\%_{\rm Latt}$ & $<5\times10^{-8}$ & $10\%$ & {\footnotesize LHCb} \\ $\mathcal{B}(B_d\to \mu^+\mu^-)$ & $1\times 10^{-10}$ & $20\% \to 5\%_{\rm Latt}$ & $<1.5\times10^{-8}$ & [?] & {\footnotesize LHCb} \\ $A_{\rm FB}(B\to K^*\mu^+\mu^-)_{q^2_0}$ & 0 & $0.05$ & $(0.2 \pm 0.2)$ & $0.05$ & {\footnotesize LHCb} \\ $B\to K\nu\bar\nu$ & $4\times 10^{-6}$ & $20\% \to 10\%_{\rm Latt}$ & $< 1.4\times 10^{-5}$ & 20\% & {\footnotesize Super-$B$} \\ \hline $|q/p|_{D-{\rm mixing}}$ & 1 & $< 10^{-3}$ & $(0.86^{+0.18}_{-0.15})$ & $0.03$ & {\footnotesize Super-$B$} \\ $\phi_D\quad$ & 0 & $< 10^{-3}$ & $−(9.6^{+8.3}_{-9.5})^\circ $ & $2^\circ$ & {\footnotesize Super-$B$} \\ \hline $\mathcal{B}(K^+\to \pi^+\nu\bar\nu)$ & $8.5 \times 10^{-11}$ & 8\% & $(1.73^{+1.15}_{-1.05})\times 10^{-10}$ & 10\% & {\footnotesize $K$ factory} \\ $\mathcal{B}(K_L \to \pi^0 \nu\bar\nu)$ & $2.6 \times 10^{-11}$ & 10\% & $<2.6 \times 10^{-8}$ & [?] & {\footnotesize $K$ factory} \\ $ R^{(e/\mu)}(K \to \pi \ell \nu)$ & $2.477 \times 10^{-5}$ & 0.04\% & $(2.498 \pm 0.014)\times 10^{-5}$ & 0.1\% & {\footnotesize $K$ factory} \\ \hline $\mathcal{B}(t \to c\,Z, \gamma)$ & ${\cal O}\left(10^{-13}\right)$ & ${\cal O}\left(10^{-13}\right)$ & $< 0.6 \times 10^{-2}$ & ${\cal O}\left(10^{-5}\right)$ & {\footnotesize LHC ($100$\,fb$^{-1}$)} \\ \hline\hline \end{tabular}\vspace*{4pt} \caption{\small Status and prospects of selected $B_{s,d}$, $D$, $K$ and $t$ observables (based on information from Ref.~\cite{Buchalla:2008jp,Grossman:2009dw,Antonelli:2009ws}). In the third column ``Latt'' refer to improvements in Lattice QCD expected in the next 5 years. In the fourth column the bounds are 90\%\,CL. The errors in the fifth column refer to 10\,fb$^{-1}$ at LHCb, 50\,ab$^{-1}$ at Super-$B$, and two years at NA62 (``$K$ factory''). In the third and fifth column the errors followed by ``\%" are relative errors, whil the others are absolute errors. For entries marked ``[?]'' we have not found a reliable estimate of the future experimental prospects. } \label{tab:future} \end{table} } While some improvements can still be expected from running experiments, in particular from CDF and D0 at Fermilab, a substantial step forward can be achieved only with the new dedicated facilities. At the LHC, which has just started to operate, the LHCb experiment is expected to collect about 10\,fb$^{-1}$ of data by 2015. Beyond that, an LHCb upgrade is planned with 10 times larger luminosity. At CERN the NA62 experiment is expected to start in 2012, with the main goal of collecting $\mathcal{O}(100)$ events of the rare decay $K^+ \to \pi^+ \nu\bar\nu$ by 2014. Dedicated rare $K$ experiments are planned also at JParc and at Fermilab. The project to upgrade KEK-B into a Super-$B$ factory, with a luminosity exceeding $5 \times 10^{35}/{\rm cm}^2s^{-1}$ has just started, and an even more ambitious super-$B$ project is currently under study in Italy. In Table~\ref{tab:future}, based on data from Ref.~\cite{Buchalla:2008jp,Grossman:2009dw,Antonelli:2009ws,TevT,Carvalho:2007yi}, we show the possible improvements for a series of particularly significant $B_{s,d}$, $D$, $K$ and $t$ observables, most of which have already been discussed in the previous sections. The future improvements refer to 10\,fb$^{-1}$ at LHCb, 50\,ab$^{-1}$ at Super-$B$, and two years of nominal data taking at NA62. The table is not comprehensive and the entries in the last two columns necessarily have significant uncertainties. However, it illustrates two important points: i) there are still several clean observables where we can expect significant improvements in the near future; ii) progress in this field requires the combined efforts of different experimental facilities. \medskip To conclude, we stress that the future of flavor physics is promising: \begin{itemize} \item The technology to collect much more flavor data exists. \item There are still several measurements which are not theory limited. \item Most well-motivated TeV-scale extensions of the Standard Model predict deviations within future experimental sensitivities and above the SM theoretical uncertainties. \end{itemize} The combination of direct discoveries at the LHC and a pattern of deviations in flavor machines is likely to solve the new physics flavor puzzle, to indirectly probe physics at a scale much higher than the TeV scale, and, perhaps to make progress in understanding the standard model flavor puzzle. {\bf Acknowledgments } We thank Kaustubh Agashe and Zoltan Ligeti for comments on the first version of this review. G.I. is supported by the EU under contract MTRN-CT-2006-035482 {\em Flavianet}. Y.N. is supported by the Israel Science Foundation (ISF) under grant No.~377/07, by the German-Israeli foundation for scientific research and development (GIF), and by the United States-Israel Binational Science Foundation (BSF), Jerusalem, Israel. G.P. is supported by the Israel Science Foundation (grant \#1087/09), EU-FP7 Marie Curie, IRG fellowship and the Peter \& Patricia Gruber Award.
\section{Introduction} \subsection{Historical overview of the theory} The formation of a stellar cusp of stars near massive black holes (MBHs) was first discussed by \citet{Pee72}, who argued that a Maxwellian distribution is not feasible because stars would be destroyed at unphysical high rates close to the MBH. \citet{Bah76} performed a more detailed calculation and showed that indeed a cusp forms, but with a shallower slope than found by Peebles. The Bahcall \& Wolf profile of $n(r)\propto r^{-7/4}$, where $n(r)$ is the density at radius $r$, has since been confirmed in many theoretical papers with different methods, including Fokker-Planck \citep[e.g.][]{Coh78, Mur91}, Monte Carlo \citep[e.g.][]{Sha78, Fre02} and $N$-body methods \citep{Pre04, Bau04a}, and can be understood (with hindsight) by simple dimensional arguments \citep[e.g.][]{Bin08, Sar06}. In a second paper, \citet{Bah77} considered the evolution of stellar systems with several masses (denoted with $M$) and showed that again the resulting steady state distribution functions $f_{M}(E) \propto E^{p_M}$ are approximate power laws, with the more massive stars having steeper slopes. They found that the stellar current into the MBH, $Q_{M} (E )$, is very small (``zero-flow solution") which leads to a specific relation between the stellar mass and the logarithmic slope of the DF\footnote{Throughout this paper the subscripts $L, H$ refer to either the light or the heavy component of a two-mass system.}, $p_M = M_L/4M_H$. In the Keplerian limit near the MBH, these DFs correspond to power-law density cusps, $n(r)\propto r^{-\alpha_M}$ with $\alpha_M = 3/2 + p_M$. In their analysis, they considered quite moderate number and mass-ratios, and concluded that the stars with the lowest masses have slopes $\alpha_L \gtrsim 3/2$, whereas the stars with the highest masses have slopes $\alpha_H \lesssim 2 $. In contrast to their work on single mass-systems, for several decades surprisingly few studies pursued the dynamics of multi-mass systems. \citet{Mur91} performed a Fokker-Planck study with a mass-spectrum, in addition to other physical processes such as stellar collisions, tidal disruptions and stellar evolution. More recently, motivated by observations of the stellar population surrounding the MBH in Galactic centre (GC), there have been several studies of mass-segregation near MBHs. The first applications to the GC were by \citet{Fre02, Fre03, Fre06}, who used H\'enon-type Monte Carlo simulations. These simulations broadly confirmed the results by \citet{Bah77}, and showed that cusps should indeed form near the MBH in our own GC. The planning of the {\it Laser Interferometer Space Antenna} (LISA) formed another motivation for the study of mass-segregation near MBHs \citep[e.g.,][]{Sig97b}. The reason for this is as follows. When compact stellar remnants spiral into MBHs of $\sim10^6 {M_{\odot}}$, they emit gravitational waves with frequencies and power such that LISA should be able to observe them out to redshifts $z\gtrsim1$. It was shown by \citet{Hop05} that stars can only spiral in successfully if they start very close ($\lesssim0.01 {\rm pc}$) to the MBH, for otherwise they are scattered away from their inspiral orbit. The implication is that inspiral rates are very sensitive to the details of the stellar distribution very close to the MBH, which is in turn dependent on the amount of mass segregation in the cusp. As a result, inspiral rates of stellar black holes (BHs) may be higher than those of white dwarfs (WDs) \citep{Hop06b}: even though for typical initial mass functions there are many more WDs than BHs in an evolved stellar population, mass-segregation drives the BHs to such orbits where they can spiral in, such that within $\sim0.01{\rm pc}$ there may in fact be more BHs than WDs. In addition to inspiral stars, LISA may also observe gravitational wave bursts from stars during a single fly-by in our {\it own} GC \citep{Rub06}. These stars are usually on very short period ($\sim1{\rm yr}$) orbits, and mass-segregation makes it again more likely to observe bursts from BHs than from other stars \citep{Hop07}. The interest in mass segregation near MBHs, driven by observation, has also led to new theoretical developments. \citet{Ale09} extended the analysis by \citet{Bah77} to a much larger parameter space, and showed that the study by \citet{Bah77} happened to be at a corner of this space where the behaviour is very stable, but that for systems where the massive stars are less numerous the evolution is very different; see \S\ref{s:duo} and Figure (\ref{f:Delta}). More specifically, they found that the most massive stars can have power law slopes that are steeper than $\alpha=2$, which was ruled out by \citet{Bah77}. This kind of ``strong mass segregation" was confirmed in $N$-body simulations by \citet{Pre10}. An analytical study of mass segregation with {\it continuous} mass functions was performed by \citet{Kes09}, who showed that, in the maximally steep limit, heavy stars can develop a cusp as steep as $n(r) \propto r^{-3}$. In this work we revisit the evolution of a cusp from a duo-mass stellar population which undergoes mass segregation \citep{Ale09}, determine the time scale for arriving at a steady state distribution, and discuss the results in terms of the GC. \section{Duo-mass systems and strong mass-segregation}\label{s:duo} \citet{Ale09} use Fokker-Planck simulations to find the the approximate steady state distribution functions (DFs) of a non-evolving spherically-distributed duo-mass stellar population around a fixed MBH. Mass is measured in units of the mass of a reference star $M_{\star}$, specific energy in units of its velocity dispersion $\epsilon_{\star}\!=\!\sigma_{\star}^{2}$, and time in units of its two-body relaxation time at the radius of influence, \begin{equation}\label{e:tstar} t_{\star}=\frac{3(2\pi\sigma_{\star}^{2})^{3/2}}{32\pi^{2}G^{2}M_{\star}^{2}n_{\star}\ln\Lambda}\,, \end{equation} where the Coulomb term is estimated as $\Lambda\!=\!M_{\bullet}/M_{\star}$. Phase space density is expressed in units of $f_{\star}\!=\! n_{\star}/(2\pi\sigma_{\star}^{2})^{3/2}$ and distance in units of the MBH radius of influence $r_{\star}\!=\! GM_{\bullet}/\sigma_{\star}^{2}$. In these units, the dimensionless specific orbital energy is defined as $x\!=\!\epsilon/\sigma_{\star}^{2}\!=\!r_{\star}/(2a)$ ($a$ is the semi-major axis), the dimensionless time is defined as $\tau\!=\! t/t_{\star}$, and the dimensionless DF of each stellar mass group as $g_{M}\!=\! f_{M}/f_{\star}$. The evolution of the dimensionless DF, $g_{M}$, is given by the time and energy dependent particle conservation equation that describes the two-body diffusion of stars in energy from a fixed unbound reservoir into the MBH sink, \begin{equation} \frac{\partial}{\partial\tau}g_{M}(x,\tau)=-x^{5/2}\frac{\partial}{\partial x}Q_{M}(x,\tau)\,.\label{e:FPeq}\end{equation} $Q_{M}$ is the energy flow integral which expresses the change in energy due to two-body scattering \citep{Bah76,Bah77}, \begin{eqnarray} Q_{M}(x) & = & \sum_{M'}MM'\int_{-\infty}^{x_{D}}\frac{\mathsf{d}x'}{\max\left(x,x'\right)^{3/2}}\times\nonumber \\ & & \left[g_{M}(x)\frac{\partial g_{M'}(x')}{\partial x'}\!-\!\frac{M'}{M}g_{M'}(x')\frac{\partial g_{M}(x)}{\partial x}\right]\,.\label{e:Qm}\end{eqnarray} Equation (\ref{e:FPeq}) is integrated in time from an arbitrary initial DF until steady state is achieved, subject to the boundary conditions that the DF falls to zero at some very high energy $x_{D}$ where the stars are destroyed, and that the unbound stars are replenished from a Maxwellian reservoir, \begin{equation} g_{M}(x\!>\! x_{D})\!=\!0\,,\quad g_{M}(x\!<\!0)\!=\! C_{M}\exp[(\sigma_{\star}^{2}/\sigma_{M}^{2})x]\,,\label{e:BCs} \end{equation} where the constant $C_{M}$ is related to $S_{M}$ (the asymptotic number density ratio of star M relative to the reference star; $S_* = 1$ by definition) by $C_{M}=\left(\sigma_{\star}/\sigma_{M}\right)^{3}S_{M}\,$. They assume violent relaxation boundary conditions such that $C_{M}\!=\! S_{M}\!=\! S_{\star}$; the steady state DFs do not depend strongly on these specific choices of boundary conditions \citep{Bah77, Ale09}. These calculations are used to explore steady state solutions for stellar distributions in both the weak and strong mass segregation regime, the nature of which they notate using the relaxational coupling parameter $\Delta$. This parameter describes the competition between the self-coupling of the heavy stars and the light-heavy coupling in terms of the mass and number ratios, \begin{equation} \Delta = {N_H M^2_H \over N_L M^2_L } \times {4 \over 3 + M_H/M_L}. \label{e:Delta} \end{equation} We reproduce some results of their simulations in Figure (\ref{f:Delta}) which shows the local logarithmic slopes of the DFs, $p_L$ and $p_H$, at $x=10$ (which corresponds to orbits with a semi major axis of $a = 0.1 $pc for $r_* = 2$ pc in the GC) as a function of $\Delta$. For comparison, they take the same mass ratios as modeled by BW77, $M_H /M_L = 1.5, 3, 10$. \begin{figure}[!h] \begin{center} \includegraphics[height=85 mm,angle=0 ]{f1.eps} \caption{The power-law indices $P_{\rm L,H}$ (thin and thick lines respectively), as derived from the logarithmic slopes of $g_{\rm L,H}$ at $x = 10$ as a function of $\Delta$ for different mass ratios. The logarithmic slopes of the DFs as calculated by \citet{Bah77} are over-plotted as open circles. The transition between the weak and strong mass segregation solutions at $\Delta \sim 1$, which is a reflection of the breakdown of the zero-flow assumption as $\Delta \rightarrow 0$, is marked. {\it Figure reproduced from \citet{Ale09} by permission of the AAS.} \label{f:Delta}} \end{center} \end{figure} In the weak segregation limit (the Bahcall-Wolf solution), $\Delta \rightarrow \infty$, which is the zero-flow ($Q_M \rightarrow 0$) limit, the heavy stars dominate the population and relax to the single mass cusp $\alpha_H = 7/4$ ($p_H=1/4$). The light stars heat by scattering against the effectively infinite reservoir of heavy stars and diffuse to lower energies, thereby settling to a flatter cusp with $\alpha_L \rightarrow 3/2$ ($p_L = M_L/4M_H \rightarrow 0$). In the strong segregation limit, $\Delta \rightarrow 0$, the light stars, which dominate the population, behave as a single mass population with $\alpha_L = 7/4$ ($p_L =1/4$). The rare heavy stars sink to the centre by dynamical friction against the effectively infinite reservoir of light stars. For low mass ratios, $M_H /M_L < 4$, where dynamical friction is less efficient, the heavy stars approximately obey the BW77 relation, $p_H= (M_H /M_L) p_L = M_H /4M_L$. For higher mass ratios, $M_H /M_L > 4$, the heavy stars approach the dynamical friction limit, $p_H \rightarrow 5/4$. \\ The present-day mass function of evolved stellar populations (coeval or continuously star forming) with a universal initial mass function, separates into two distinct mass scales, $\sim 1 M_{\odot}$ of main sequence stars, dwarfs and neutron stars, and $\sim 10 M_{\odot}$ of BHs. In this sense such systems can be approximated by a duo-mass system. Furthermore, \citet{Ale09} show that $\Delta < 0.1$ so that this system should be strongly mass segregated. In the next section we calculate the time scale for such a system to form a mass segregated steady state cusp. \section{Time-dependence of cusp formation}\label{s:time} The amount of time it takes to form a stellar cusp around a MBH in a galactic nucleus is important for many reasons including, but not limited to, event rates of stellar collisions, tidal disruptions of stars and inspirals of compact objects with the emission of gravitational waves. In particular, if this time is greater than a Hubble time, event rates will be lower than previously predicted. To determine this time scale we use Fokker-Planck simulations as described in \S\ref{s:duo} with initial DFs $g_M(x)\propto x^{-0.5}$. We find that steady state is reached in $\tau = 0.2-0.3$; see Figure (\ref{f:ss}). Our result is in agreement with \citet{Pre10} who find, using both Fokker-Planck and $N$-body simulations, that steady state is reached in $t \sim (0.1 - 0.2) ~ T_{\rm rlx}(r_h)$, where $T_{\rm rlx}(r_h)$ is the value of the relaxation time at the radius of influence of a MBH, and roughly equivalent to $t_{\star}$ in Equation (\ref{e:tstar}). Values of the relaxation time for MBHs with $M_{\bullet} \lesssim 6 \times 10^6 M_{\odot}$ vary in the literature, but are generally constrained to within $0.5 - 3 \times 10^{10} {\rm yr}$. Our results imply that quasi-steady, mass segregated stellar cusps should be common around MBHs in galactic nuclei for this mass regime. Hence, assuming GC parameters that we observe today \citep[see][]{Mer09b}, a cusp should form within \begin{eqnarray} t_{\rm cusp} & = & 0.2 -0.3~ t_{\star} \nonumber\\ & \approxeq & 0.5 - 6 \times 10^9 {\rm yr} . \label{e:tcusp} \end{eqnarray} \begin{figure}[!h] \begin{center} \includegraphics[height=120 mm,angle=270 ]{f2.eps} \caption{Evolution of the distribution function, shown at dimensionless times $\tau=\{0.1, 0.2, 0.3\}$ with increasing thickness for increasing time. Steady state is reached in about $\tau=0.2-0.3$. Initially, all DFs were $g(x)\propto x^{-0.5}$. Interestingly, for short times the distribution of white dwarfs (WD), main sequence stars (MS) and neutron stars (NS) grows steeper than the final distribution which is approximately energy independent. The reason is that it takes time for the stellar black holes (BH) to grow a steep cusp and become the dominant species. \label{f:ss}} \end{center} \end{figure} \section{Applications} \subsection{Galactic centre cusp} Several studies \citep[e.g.][]{Ale99a, Gen03a, Sch07} have shown that indeed a cusp of stars is present in the GC. However, recently it has become clear that even though the {\it young} stars do have a power law profile, there is a dearth of giant stars within $\sim0.1{\rm pc}$ \citep{Buc09, Do09, Bar10}. Since the O/B stars are too young to have relaxed by the two-body relaxation process discussed here, the conclusion is that currently there is no observational support for cusp formation or mass-segregation in the GC. The reason for the absence of a cusp is unclear. One possibility is that physics not considered here depletes the tightly bound orbits of old stars, for example as a result of stellar collisions \citep[e.g.,][]{Dal09}. Another option was considered by \citet{Mer09b}, who suggested that relaxational physics {\it does} in fact capture the main mechanism for cusp formation, but that the age of the GC is too short compared to the time scale for a cusp to have formed today. His estimate of the relaxation time is $\gtrsim 2 - 3 \times 10^{10} ~{\rm yr}$, and in his Fokker-Planck simulations, \citet{Mer09b} does not find the formation of a relaxed cusp within that time. This is in contrast to the result we found in \S\ref{s:time}, where a cusp forms in $0.2-0.3 ~ t_{\star}$, similar to the findings by \citet{Pre10}. Thus, even for a relaxation time of $30$ Gyr, the central pc would form a cusp in less than a Hubble time, {\it starting from the current situation}. The reason for this discrepancy in the time scale for reaching a steady state solution is not yet clear, but may be related to the back reaction of the BHs on the light main sequence stars \citep{Pre10}. We note that our findings do not imply that the GC is necessarily relaxed and formed a cusp: it is possible that in the past the density at the radius of influence was lower, that it has taken a Hubble time to contract\footnote{Our treatment of the Fokker-Planck equations does not evolve the stars that are unbound to the MBH, such that we cannot follow the contraction of the cusp. For a study that includes self-consistent mass-segregation and evolution of the ambient cluster of stars, see \citet{Pre10}.} to the current configuration, and that the remaining time to form a cusp is less than a Hubble time \citep{Mer09b}. However, this argument appears to be more attractive when the remaining time for cusp formation exceeds a Hubble time. \subsection{S-stars} Relaxational dynamics cannot account for the presence of a cluster of B-stars known as the``S-stars" in the inner $\sim0.03 {\rm pc}$ near the MBH in the GC \citep{Sch02, Ghe05, Eis05, Gil09}, since the time scales are too long for young stars (with ages between $\sim 20 -100$ Myr) to diffuse to such distances. There have been many suggestions of how the S-stars have reached their current orbits. Currently the most promising scenarios seem to be those where a binary star is tidally disrupted by the MBH \citep{Hil88, YuQ03}. Such binaries may originate very far from the MBH, where they are perhaps driven to eccentric orbits by triaxiality or massive perturbers \citep{Per07}. Alternatively, binaries which form in eccentric accretion discs can be pumped to very high eccentricities by a gravitational instability \citep{Mad09}. In either case, a B-star may end up on a tightly bound orbit, but the orbit will be very eccentric ($e\!\sim\!0.99$). Post-capture dynamics must then account for the mapping of the initial, very high eccentricity distribution to the current distribution in the GC. The mechanism responsible for changing the eccentricities is most likely resonant relaxation \citep[RR;][]{Rau96}, a secular process that very efficiently randomizes the angular momenta of orbits if the precession time is much longer than the orbital time-scale. For the S-stars, the resonant relaxation time is of order \begin{equation} T_{\rm RR} \sim 100\, {\rm Myr} {M_{\odot}\over M_{\star}}, \end{equation} where $M_{\star}$ is the typical mass of stars. For $M_{\star}=10M_{\odot}$, appropriate for mass segregated BHs, this time-scale is short enough to randomize the orbits \citep{Hop06, Lev07}. Post-capture evolution was studied in detail by \citet{Per09} with the use of $N$-body simulations. These simulations do not include general relativistic precession, which is potentially important for the highly eccentric orbits that are considered here, but they clearly show that the distribution evolves on time-scales of the order $T_{\rm RR}$. For $M_{\star} = 1 M_{\odot}$ however, this time scale is too large for the S-star orbits to have evolved from high eccentricities to their currently observed orbits. Only if there are mass-segregated BHs within the S-star radii can RR explain their orbital parameters. \subsection{Gravitational waves} As stated in the introduction, most compact remnants that spiral in successfully onto MBHs to become detectable gravitational wave sources originate from distances very close ($\lesssim0.01 {\rm pc}$) to the MBH. It is therefore important to what extent such orbits are populated, and mass-segregation helps to increase this number. In spite of this, even if there is no cusp in the GC, this would not necessarily imply that other, similar galaxies do not have cusps either. If, as we find, the time for cusp formation from the current state is less than a Hubble time, then one would expect that most galaxies with MBHs similar to the GC did in fact form cusps, even if our Galaxy did not. Furthermore, most inspiral sources originate from lower mass MBHs \citep{Hop09, Gai09}; such galaxies have even shorter relaxation times, and are thus more likely to have formed cusps. The predicted number of detectable LISA sources will therefore not depend strongly on the question of whether a cusp is present or not in the GC. Conclusions regarding the absence of a gravitational wave background from fly-bys \citep{Too09} will likewise not be strongly affected.
\section{Introduction} Experiments on neutrino oscillations, which measure differences of squared masses and mixing angles \cite{Altarelli:2004, Altarelli:2009, Mohapatra:2006, Mohapatra:2007, Grimus:2006, Gonzalez-Garcia:2008} have established that neutrinos have a mass. We refer in particular to ref. \cite{Altarelli:2004} for an introduction to the subject, the main results, the basic formalism and all definitions and notations. Two distinct oscillation frequencies have been first measured in solar \cite{Super-Kamiokande:2006, Super-Kamiokande:2008, SNO:2009} and atmospheric \cite{Super-Kamiokande:2006a, Super-Kamiokande:2006b} neutrino oscillations and later confirmed by experiments on earth, like KamLAND \cite{KamLAND:2008}, K2K \cite{K2K:2006}, MINOS \cite{Kafka:2010zz} and OPERA \cite{Agafonova:2010dc}. A signal corresponding to a third mass difference was claimed by the LSND experiment \cite{LSND:1996, LSND:1998, LSND:1998a} but not confirmed by KARMEN \cite{KARMEN:2002} and recently by MiniBooNE \cite{MiniBooNE:2009, MiniBooNE:2009a}. Two well separated differences need at least three different neutrino mass eigenstates involved in oscillations. Actually the three known neutrino species can be sufficient. At least two $\nu$'s must be massive while, in principle, the third one could still be massless. In the following we will assume the simplest picture with three active neutrinos, no sterile neutrinos and CPT invariance. The mass eigenstates involved in solar oscillations are $m_1$ and $m_2$ and, by definition, $|m_2|> |m_1|$, so that $\Delta m^2_{sun}=\Delta m^2_{21}=|m_2|^2-|m_1|^2>0$. The atmospheric neutrino oscillations involve $m_3$: $\Delta m^2_{atm}=|\Delta m^2_{31}|$ with $\Delta m^2_{31}=|m_3|^2-|m_1|^2$ either positive (normal hierarchy) or negative (inverse hierarchy). The present data \cite{Strumia:2006, Gonzalez-Garcia:2008a, Bandyopadhyay:2008, Fogli:2008, Fogli:2008a, Schwetz:2008, Maltoni:2008} are compatible with both cases. The degenerate spectrum occurs when the average absolute value of the masses is much larger than all mass squared differences: $|m_i|^2 >> |\Delta m^2_{hk}|$. With the standard set of notations and definitions \cite{Altarelli:2004} the present data are summarised in Table 1. \begin{table}[h] \begin{center} \begin{tabular}{|c|c|c|} \hline Quantity & ref. \cite{Fogli:2008, Fogli:2008a} & ref. \cite{Schwetz:2008, Maltoni:2008} \\ \hline $\Delta m^2_{sun}~(10^{-5}~{\rm eV}^2)$ &$7.67^{+0.16}_{-0.19}$ & $7.65^{+0.23}_{-0.20}$ \\ $\Delta m^2_{atm}~(10^{-3}~{\rm eV}^2)$ &$2.39^{+0.11}_{-0.08}$ & $2.40^{+0.12}_{-0.11}$ \\ $\sin^2\theta_{12}$ &$0.312^{+0.019}_{-0.018}$ & $0.304^{+0.022}_{-0.016}$ \\ $\sin^2\theta_{23}$ &$0.466^{+0.073}_{-0.058}$ & $0.50^{+0.07}_{-0.06}$ \\ $\sin^2\theta_{13}$ &$0.016\pm0.010$ &$0.010^{+0.016}_{-0.011}$ \\ \hline \end{tabular} \end{center} \caption{\label{tab:data} Fits to neutrino oscillation data.} \end{table} Oscillation experiments do not provide information about either the absolute neutrino mass scale or the Dirac/Majorana nature of neutrinos. Limits on the mass scale are obtained \cite{Altarelli:2004} from the endpoint of the tritium beta decay spectrum, from cosmology (see, for example \cite{Lesgourgues:2006}) and from neutrinoless double beta decay ($0\nu \beta \beta$) (for a recent review, see, for example \cite {Avignone:2008}). From tritium we have an absolute upper limit of 2.2 eV (at 95\% C.L.) on the mass of electron antineutrino \cite{Kraus:2005}, which, combined with the observed oscillation frequencies under the assumption of three CPT-invariant light neutrinos, represents also an upper bound on the masses of the other active neutrinos. Complementary information on the sum of neutrino masses is also provided by the galaxy power spectrum combined with measurements of the cosmic microwave background anisotropies. According to recent analyses of the most reliable data \cite{Fogli:2008b} $\sum_i \vert m_i\vert < 0.60\div 0.75$ eV (at 95\% C.L.) depending on the retained data (the numbers for the sum have to be divided by 3 in order to obtain a limit on the mass of each neutrino). The discovery of $0\nu \beta \beta$ decay would be very important because it would establish lepton number violation and the Majorana nature of $\nu$'s, and provide direct information on the absolute scale of neutrino masses. The present limit from $0\nu \beta \beta$ (with large ambiguities from nuclear matrix elements) is about $\vert m_{ee}\vert < (0.3\div 0.8)$ eV \cite {Avignone:2008, Fogli:2008b} (see eq. (\ref{3nu1gen})). After KamLAND \cite{KamLAND:2008}, SNO \cite{SNO:2009} and the upper limits on the absolute value of neutrino masses not too much hierarchy in the spectrum of neutrinos is indicated by experiments: \begin{eqnarray} r = \Delta m_{sol}^2/\Delta m_{atm}^2 \sim 1/30~~~. \label{r} \end{eqnarray} Precisely $r=0.032^{+0.006}_{-0.005}$ at $3\sigma$'s \cite{Fogli:2008, Fogli:2008a, Schwetz:2008, Maltoni:2008}. Thus, for a hierarchical spectrum, $m_2/m_3 \sim \sqrt{r} \sim 0.2$, which is comparable to the Cabibbo angle $\lambda_C \sim 0.22$ or to its leptonic analogue $\sqrt{m_{\mu}/m_{\tau}} \sim 0.24$. This suggests that the same hierarchy parameter (raised to powers with $\mathcal{O}(1)$ exponents) may apply for quark, charged lepton and neutrino mass matrices. This in turn indicates that, in the absence of some special dynamical reason, we do not expect quantities like $\theta_{13}$ or the deviation of $\theta_{23}$ from its maximal value to be too small. Indeed it would be very important to know how small the mixing angle $\theta_{13}$ is and how close to maximal $\theta_{23}$ is. Given that neutrino masses are certainly extremely small, it is really difficult from the theory point of view to avoid the conclusion that the lepton number $L$ conservation is probably violated and that $\nu$'s are Majorana fermions. In this case the smallness of neutrino masses can be naturally explained as inversely proportional to the large scale where $L$ conservation is violated. If neutrinos are Majorana particles, their masses arise from the generic dimension-five non renormalizable operator of the form \cite{Weinberg:1979}: \begin{equation} O_5=\frac{(H l)^T_i \eta_{ij} (H l)_j}{M}+~h.c.~~~, \label{O5} \end{equation} with $H$ being the ordinary Higgs doublet, $l_i$ the SU(2) lepton doublets, $\eta$ a matrix in flavor space, $M$ a large scale of mass and a charge conjugation matrix $C$ between the lepton fields is understood. For $\eta_{ij}\approx \mathcal{O}(1)$, neutrino masses generated by $O_5$ are of the order $m_{\nu}\approx v^2/M$ where $v\sim {\rm O}(100~{\rm GeV})$ is the vacuum expectation value of the ordinary Higgs. A particular realization of this effective mass operator is given by the see-saw mechanism \cite{Minkowski:1977, Yanagida:1979, Gell-Mann:1979, Glashow:1980, Mohapatra:1980} , where $M$ derives from the exchange of heavy neutral objects of weak isospin 0 or 1. In the simplest case the exchanged particle is the right-handed (RH) neutrino $\nu^c$ (a gauge singlet fermion here described through its charge conjugate field), and the resulting neutrino mass matrix reads (type I see-saw ) \cite{Altarelli:2004}: \begin{equation} m_{\nu}=m_D^T M^{-1}m_D~~~, \end{equation} where $m_D$ and $M$ denote the Dirac neutrino mass matrix (defined as $ {\nu^c}^T m_D \nu$) and the Majorana mass matrix of $\nu^c$ (defined as $ {\nu^c}^T M \nu^c$), respectively. As one sees, the light neutrino masses are quadratic in the Dirac masses and inversely proportional to the large Majorana mass. For $m_{\nu}\approx \sqrt{\Delta m^2_{atm}}\approx 0.05$ eV and $m_{\nu}\approx m_D^2/M$ with $m_D\approx v \approx 200$~GeV we find $M\approx 10^{15}$~GeV which indeed is an impressive indication that the scale for lepton number violation is close to the grand unified scale $M_{GUT}$. Thus probably neutrino masses are a probe into the physics near $M_{GUT}$. This argument, in our opinion, strongly discourages models where neutrino masses are generated near the weak scale and are suppressed by some special mechanism. Oscillation experiments cannot distinguish between Dirac and Majorana neutrinos. The detection of neutrino-less double beta decay would provide direct evidence of $L$ non conservation, and the Majorana nature of neutrinos. It would also offer a way to possibly disentangle the 3 cases of degenerate, normal or inverse hierachy neutrino spectrum. The quantity which is bound by experiments on $0\nu \beta \beta$ is the 11 entry of the $\nu$ mass matrix, which in general, from $m_{\nu}=U^* m_{diag} U^\dagger$, is given by : \begin{eqnarray} \vert m_{ee}\vert~=\vert(1-s^2_{13})~(m_1 c^2_{12}~+~m_2 s^2_{12})+m_3 e^{2 i\phi} s^2_{13}\vert~~~, \label{3nu1gen} \end{eqnarray} where $U\equiv U_{PMNS}$ is the mixing matrix, $m_{1,2}$ are complex masses (including Majorana phases) while $m_3$ can be taken as real and positive and $\phi$ is the $U_{PMNS}$ phase measurable from CP violation in oscillation experiments. Starting from this general formula it is simple to derive the bounds for degenerate, inverse hierarchy or normal hierarchy mass patterns shown in Fig. 1 \cite{Feruglio:2002}. In the next few years a new generation of experiments will reach a larger sensitivity on $0\nu \beta \beta$ by about an order of magnitude. If these experiments will observe a signal this will be compatible with both type of neutrino mass ordering, if not, then the normal hierarchy case remains a possibility. Establishing that $L$ is violated in particle interactions would also strongly support the possibility that the observed baryon asymmetry is generated via leptogenesis, through the out-of-equilibrium, CP and $L$ violating decays of the heavy RH neutrinos (see Sect. 10). \begin{figure} \centering \includegraphics [width=10.0 cm]{nuless.eps} \caption{A plot \cite{Feruglio:2002} of $m_{ee}$ in eV, the quantity measured in neutrino-less double beta decay, given in eq. (\ref{3nu1gen}), versus the lightest neutrino mass $m_1$, also in eV. The upper (lower) band is for inverse (normal) hierarchy.} \end{figure} Neutrino mixing is important because it could in principle provide new clues for the understanding of the flavor problem. Even more so since neutrino mixing angles show a pattern that is completely different than that of quark mixing: for quarks all mixing angles are small, for neutrinos two angles are large (one is even compatible with the maximal value) and only the third one is small. For building up theoretical models of neutrino mixing one must guess which features of the data are really relevant in order to identify the basic principles for the formulation of the model. In particular, it is an experimental fact \cite{Strumia:2006, Gonzalez-Garcia:2008a, Bandyopadhyay:2008, Fogli:2008, Fogli:2008a, Schwetz:2008, Maltoni:2008} that within measurement errors the observed neutrino mixing matrix \cite{Altarelli:2004} is compatible with the so called Tri-Bimaximal (TB) form in eq. (\ref{2}) \cite{Harrison:2002, Harrison:2002a, Harrison:2003, Harrison:2004}. The best measured neutrino mixing angle $\theta_{12}$ is just about 1$\sigma$ below the TB value $\tan^2{\theta_{12}}=1/2$, while the other two angles are well inside the 1$\sigma$ interval (see table \ref{tab:data}). Thus, one possibility is that one takes this coincidence seriously and only considers models where TB mixing is automatically a good first approximation. Alternatively one can assume that the agreement of the data with TB mixing is accidental. Indeed there are many models that fit the data and yet TB mixing does not play any role in their architecture. For example, in ref. \cite{Albright:2008} there is a list of Grand Unified SO(10) models with parameters that can be fitted to the neutrino mixing angles leading to a good agreement with the data although most of these models have no built-in relation with TB mixing (see also \cite{Bertolini:2006}). Another class of examples is found in ref. \cite{Plentinger:2008}. Clearly, for this type of models, in most cases different mixing angles could also be accommodated by simply varying the fitted values of the parameters. If instead we assume that TB mixing has a real physical meaning, then we are led to consider models that naturally produce TB mixing in first approximation and only a very special dynamics can lead to this peculiar mixing matrix. Discrete non abelian groups (for an introduction see, for example, \cite{Frampton:1995,Ishimori:2010au}) naturally emerge as suitable flavor symmetries. In fact the TB mixing matrix immediately suggests rotations by fixed, discrete angles. It has been found that a broken flavor symmetry based on the discrete group $A_4$ (the group of even permutations of 4 elements, which can be seen as the invariance group of a rigid regular tetrahedron) appears to be particularly suitable to reproduce this specific mixing pattern in leading order (LO). A non exhaustive list of papers that discuss the application of $A_4$ to neutrino mixing is given by \cite{Ma:2001, Ma:2002, Babu:2003, Hirsch:2004, Ma:2004, Ma:2004a, Chen:2005, Altarelli:2005, Ma:2005, Hirsch:2005, Babu:2005, Ma:2005a, Zee:2005, Ma:2006, He:2006, Adhikary:2006, Altarelli:2006, Lavoura:2006, Ma:2007, Hirsch:2007, Altarelli:2007, Yin:2007, Bazzocchi:2008, Bazzocchi:2008a, Honda:2008, Brahmachari:2008, Adhikary:2008, Hirsch:2008, Frampton:2008, Csaki:2008, Altarelli:2008, Morisi:2009, Lin:2009, Lin:2009a, Altarelli:2009a, Ma:2005b, Ma:2006a, Ma:2006b, Morisi:2007, Grimus:2008, Ciafaloni:2009, Bazzocchi:2009, Bazzocchi:2008b, delAguila:2010,Kadosh:2010rm,Antusch:2010es}. The choice of this particular discrete group is not unique and, for example, other solutions based on alternative discrete flavor symmetries (for example, $T$' \cite{Frampton:1995, Aranda:2000, Aranda:2000a, Carr:2000, Aranda:2007, Frampton:2007, Frampton:2009, Ding:2009,Feruglio:2007, Chen:2007}, $S_4$ \cite{Mohapatra:2004, Hagedorn:2006, Cai:2006, Ma:2007a, Bazzocchi:2008c, Ishimori:2009, Bazzocchi:2009a, Bazzocchi:2009b, Meloni:2009, Dutta:2009, Dutta:2009a,Ding:2010,Morisi:2010,Hagedorn:2010th,Ishimori:2010xk}, $\Delta(27) $ \cite{deMedeirosVarzielas:2007, Ma:2007b, Grimus:2007, Luhn:2007, Bazzocchi:2009c, Ding:2010} and other groups \cite{Everett:2009, Luhn:2007a, King:2009, King:2009a, Luhn:2007b}) or continuous flavor symmetries \cite{King:2005, King:2006, deMedeirosVarzielas:2006, deMedeirosVarzielas:2007a, Adulpravitchai:2009c, Berger:2009} have also been considered (for other approaches to TB mixing see \cite{Xing:2002, Matias:2005, Luo:2005, Grimus:2005, Koide:2007, Grimus:2009,Babu:2010bx}), but the $A_4$ models have a particularly economical and attractive structure, e.g. in terms of group representations and of field content. In most of the models $A_4$ is accompanied by additional flavor symmetries, either discrete like $Z_N$ or continuous like U(1), which are necessary to eliminate unwanted couplings, to ensure the needed vacuum alignment and to reproduce the observed mass hierarchies. Given the set of flavor symmetries and having specified the field content, the non leading corrections to TB mixing arising from higher order effects can be evaluated in a well defined expansion. In the absence of specific dynamical tricks, in a generic model, all the three mixing angles receive corrections of the same order of magnitude. Since the experimentally allowed departures of $\theta_{12}$ from the TB value $\sin^2{\theta_{12}}=1/3$ are small, at most of $\mathcal{O}(\lambda_C^2)$ , with $\lambda_C$ the Cabibbo angle, it follows that both $\theta_{13}$ and the deviation of $\theta_{23}$ from the maximal value are typically expected in these models to also be at most of $\mathcal{O}(\lambda_C^2)$ \footnote{By $\mathcal{O}(\lambda_C^2)$ we mean numerically of order $\lambda_C^2$. As $\lambda_C \sim 0.22$ a linear term in $\lambda_C$ with a smallish coefficient can easily be $\mathcal{O}(\lambda_C^2)$}. A value of $\theta_{13} \sim \mathcal{O}(\lambda_C^2)$ is within the sensitivity of the experiments which are now in preparation and will take data in the near future. Going back to the possibility that the agreement of the data with TB mixing is accidental, we observe that the present data do not exclude a value for $\theta_{13}$, i.e. $\theta_{13} \sim \mathcal{O}(\lambda_C)$, larger than generally implied by models with approximate TB mixing. In fact, recent analysis of the available data lead to $\sin^2{\theta_{13}}=0.016\pm0.010$ at 1$\sigma$ \cite{Fogli:2008, Fogli:2008a}, $\sin^2{\theta_{13}}=0.010^{+0.016}_{-0.011}$ at 1$\sigma$ \cite{Schwetz:2008, Maltoni:2008}, $\sin^2{\theta_{13}}=0.014^{+0.013}_{-0.011}$ at 1$\sigma$ \cite{GonzalezGarcia:2010} and $\sin^2{\theta_{13}}=0.010^{+0.013}_{-0.009}$ at 1$\sigma$ \cite{GonzalezGarcia:2010}, which are compatible with both options. If experimentally it is found that $\theta_{13}$ is near its present upper bound, this could be interpreted as an indication that the agreement with the TB mixing is accidental. In fact a different empirical observation is that $\theta_{12}+\lambda_C\sim \pi/4$, a relation known as quark-lepton complementarity \cite{Raidal:2004, Minakata:2004}, or similarly $\theta_{12}+\sqrt{m_\mu/m_\tau} \sim \pi/4$. No compelling model leading, without parameter fixing, to the exact complementarity relation has been produced so far. Probably the exact complementarity relation is to be replaced with something like $\theta_{12}+\mathcal{O}(\lambda_C)\sim \pi/4$ or $\theta_{12}+\mathcal{O}(\sqrt{m_\mu/m_\tau})\sim \pi/4$ (which we could call "weak" complementarity \cite{Altarelli:2004a, Frampton:2005, Ferrandis:2005, Kang:2005, Minakata:2005, Li:2005, Cheung:2005, Xing:2005, Datta:2005, Ohlsson:2005, Antusch:2005, Lindner:2005, King:2005a, Dighe:2006, Schmidt:2006, Chauhan:2007, Hochmuth:2007, Plentinger:2007, Plentinger:2008a, Altarelli:2009b}. If we take any of these complementarity relations as a serious hint then a scheme would be relevant where Bimaximal (BM) mixing, instead of TB mixing, is the correct first approximation, modified by terms of $\mathcal{O}(\lambda_C)$. A comparison of the TB or BM mixing values with the data on $\sin^2{\theta_{12}}$ is shown in Fig. (\ref{TBBM}). \begin{figure} \centering \includegraphics [width=10.0 cm]{TBvsBM.eps} \caption{The values of $\sin^2{\theta_{12}}$ for TB o BM mixing are compared with the data} \label{TBBM} \end{figure} A very special dynamics is also needed for BM mixing and again discrete symmetry groups offer possible solutions. For example, a model \cite{Altarelli:2009b} based on $S_4$, the permutation group of 4 elements, naturally leads to BM mixing in LO. This model is built in such a way that the dominant corrections to the BM mixing only arise from the charged lepton sector at Next-to-the-Leading-Order (NLO) and naturally inherit $\lambda_C$ as the relevant expansion parameter. As a result the mixing angles deviate from the BM values by terms of $\mathcal{O}(\lambda_C)$ (at most), and weak complementarity holds. A crucial feature of this particular model is that only $\theta_{12}$ and $\theta_{13}$ are corrected by terms of $\mathcal{O}(\lambda_C)$ while $\theta_{23}$ is unchanged at this order (which is essential to make the model agree with the present data). Other types of LO approximations for the lepton mixing pattern have been suggested. For instance a viable first approximation of the solar mixing angle is also $\theta_{12}=\tan^{-1}(1/\varphi)$ where $\varphi=(1+\sqrt{5})/2$ is the golden ratio \cite{Kajiyama:2007}. This leads to $\sin^2\theta_{12}=1/(1+\varphi^2)\approx 0.276$, not far from the allowed range. Another possible connection with the golden ratio has been proposed in ref. \cite{Rodejohann:2009}. In this case $\cos\theta_{12} = \varphi/2$, or $\sin^2\theta_{12}= 1/4 (3 - \varphi) \approx 0.345$. There have been attempts to reproduce these values by exploiting flavor symmetries of icosahedral type \cite{Everett:2009}, for the first possibility, or of dihedral type \cite{Adulpravitchai:2009} for the second case. Thus discrete flavor symmetries may play an important role in models of neutrino mixing. In particular this is the case if some special patterns indicated by the data as possible first approximations, like TB or BM mixing or others, are indeed physically relevant. A list of the simplest discrete groups that have been considered for neutrino mixing, with some of their properties, is shown in Table 2. In the present review we will discuss the formalism and the physics of a non exhaustive list of models of neutrino mixing based on discrete symmetries. \begin{table}[t] \begin{center} \begin{tabular}{|l|c|c|c|c|} \hline \textbf{Group} & d & Irr. Repr.'s&Presentation&Ref.'s\\ \hline $D_3\sim S_3$ & 6 & 1, $1'$, 2& $A^3=B^2=(AB)^2=1$&[i]\\ \hline $D_4$ & 8&$1_1,...1_4, 2$& $A^4=B^2=(AB)^2=1$&[ii]\\ \hline $D_7$ & 14&1, $1'$, $2$, $2'$, $2''$& $A^7=B^2=(AB)^2=1$&[iii]\\ \hline $A_4$ & 12& 1, $1'$, $1''$, 3&$A^3=B^2=(AB)^3=1$&[iv]\\ \hline $A_5 \sim PSL_2(5)$ & 60& 1, 3, $3'$, 4, 5&$A^3=B^2=(BA)^5=1$&[v]\\ \hline $T'$ & 24& 1, $1'$, $1''$, $2$, $2'$, $2''$, 3&$A^3=(AB)^3=R^2=1,~ B^2=R$&[vi]\\ \hline $S_4$ & 24 & 1, $1'$, 2, 3, $3'$&$BM: A^4=B^2=(AB)^3=1$&\\ $$ & &&$TB: A^3=B^4=(BA^2)^2=1$&[vii]\\ \hline $\Delta(27) \sim Z_3 ~\rtimes~ Z_3$ &27& $1_1,...1_9, 3, \overline{3}$&&[viii]\\ \hline $PSL_2(7)$ &168&$ 1, 3,\overline{3}, 6, 7, 8$&$A^3=B^2=(BA)^7=(B^{-1}A^{-1}BA)^4=1$&[ix]\\ \hline $T_7 \sim Z_7~\rtimes~Z_3$ & 21&$ 1, 1', \overline{1'}, 3, \overline{3}$&$A^7=B^3=1,~AB=BA^4$&[x]\\ \hline \end{tabular} \caption{Some small discrete groups used for model building. [i]\cite{Kubo:2003, Kubo:2004, Kubo:2006, Chen:2004, Lavoura:2005, Dermisek:2005, Caravaglios:2005, Caravaglios:2005a, Grimus:2006a, Koide:2006, Teshima:2006, Haba:2006, Tanimoto:2006, Koide:2006a, Morisi:2006, Picariello:2006, Mohapatra:2006a, Mohapatra:2006b, Kaneko:2007, Koide:2007a, Chen:2008, Feruglio:2007a}; [ii]\cite{Grimus:2004, Adulpravitchai:2009a}; [iii]\cite{Blum:2004, Blum:2008};[iv]\cite{Ma:2001, Ma:2002, Babu:2003, Hirsch:2004, Ma:2004, Ma:2004a, Chen:2005, Altarelli:2005, Ma:2005, Hirsch:2005, Babu:2005, Ma:2005a, Zee:2005, Ma:2006, He:2006, Adhikary:2006, Altarelli:2006, Lavoura:2006, Ma:2007, Hirsch:2007, Altarelli:2007, Yin:2007, Bazzocchi:2008, Bazzocchi:2008a, Honda:2008, Brahmachari:2008, Adhikary:2008, Hirsch:2008, Frampton:2008, Csaki:2008, Altarelli:2008, Morisi:2009, Lin:2009, Lin:2009a, Altarelli:2009a, Ma:2005b, Ma:2006a, Ma:2006b, Morisi:2007, Grimus:2008, Ciafaloni:2009, Bazzocchi:2009, Bazzocchi:2008b, delAguila:2010,Kadosh:2010rm,Antusch:2010es}; [v]\cite{Everett:2009}; [vi]\cite{Frampton:1995, Aranda:2000, Aranda:2000a, Carr:2000, Aranda:2007, Frampton:2007, Frampton:2009, Ding:2009,Feruglio:2007, Chen:2007}; [vii]\cite{Mohapatra:2004, Hagedorn:2006, Cai:2006, Zhang:2007,Ma:2007a, Bazzocchi:2008c, Ishimori:2009, Bazzocchi:2009a, Bazzocchi:2009b, Meloni:2009, Dutta:2009, Dutta:2009a, Ding:2010, Morisi:2010, Hagedorn:2010th,Ishimori:2010xk}; [viii]\cite{deMedeirosVarzielas:2007, Ma:2007b, Grimus:2007, Luhn:2007, Bazzocchi:2009c}; [ix]\cite{Luhn:2007a, King:2009, King:2009a}; [x]\cite{Luhn:2007b}.} \label{groups} \end{center} \end{table} \section{Special patterns of neutrino mixing} Given the PNMS mixing matrix $U$ (we refer the reader to ref. \cite{Altarelli:2004} for its general definition and parametrisation), the general form of the neutrino mass matrix, in terms of the (complex \footnote{We absorb the Majorana phases in the mass eigenvalues $m_i$, rather than in the mixing matrix $U$. The dependence on these phases drops in neutrino oscillations.}) mass eigenvalues $m_1, m_2, m_3$, in the basis where charged leptons are diagonal, is given by: \begin{equation} m_{\nu}=U^* {\rm diag}(m_1,m_2,m_3) U^\dagger~~~. \label{numass} \end{equation} We will present here a number of particularly relevant forms of $U$ and $m_\nu$ that will be important in the following. We start by the most general mass matrix that corresponds to $\theta_{13}=0$ and $\theta_{23}$ maximal, that is to $U$ given by (in a particular phase convention): \begin{equation} U= \left(\matrix{ c_{12}&s_{12}&0\cr -\displaystyle\frac{s_{12}}{\sqrt 2}&\displaystyle\frac{c_{12}}{\sqrt 2}&-\displaystyle\frac{1}{\sqrt 2}\cr -\displaystyle\frac{s_{12}}{\sqrt 2}&\displaystyle\frac{c_{12}}{\sqrt 2}&\displaystyle\frac{1}{\sqrt 2}}\right)~~~, \label{2.1} \end{equation} with $c_{12}\equiv \cos{\theta_{12}}$ and $s_{12}\equiv \sin{\theta_{12}}$. By applying eq. (\ref{numass}) we obtain a matrix of the form \cite{Fukuyama:1997, Mohapatra:1999, Ma:2001a, Lam:2001, Kitabayashi:2003, Grimus:2003, Koide:2004, Ghosal:2003, Grimus:2005a, deGouvea:2004, Mohapatra:2005, Kitabayashi:2005, Mohapatra:2005a, Mohapatra:2005b, Mohapatra:2005c, Ahn:2006}: \begin{equation} m=\left(\matrix{ x&y&y\cr y&z&w\cr y&w&z}\right), \label{gl} \end{equation} with complex coefficients $x$, $y$, $z$ and $w$. This matrix is the most general one that is symmetric under 2-3 (or $\mu - \tau$) exchange or: \begin{eqnarray} m_\nu=A_{23}m_\nu A_{23}~~\label{inv.1}~~~, \label{mmutau} \end{eqnarray} where $A_{23}$ is given by: \begin{equation} A_{23}=\left( \begin{array}{ccc} 1&0&0\\ 0&0&1\\ 0&1&0 \end{array} \right)~~~. \label{Amutau} \end{equation} The solar mixing angle $\theta_{12}$ is given by \begin{eqnarray} \sin^2 2 \theta_{12}&=&\displaystyle\frac{8\vert x^* y+y^*(w+z)\vert^2}{8\vert x^* y+y^*(w+z)\vert^2+(\vert w+z\vert^2-\vert x\vert^2)^2}\nonumber\\ &=&\displaystyle\frac{8 y^2}{(x-w-z)^2+8 y^2}~~~ \label{teta12} \end{eqnarray} where the second equality applies to real parameters. Since $\theta_{13}=0$ there is no CP violation in neutrino oscillations, and the only physical phases are the Majorana ones, accounted for by the general case of complex parameters. We restrict here our consideration to real parameters. There are four of them in eq. (\ref{gl}) which correspond to the three mass eigenvalues and one remaining mixing angle, $\theta_{12}$. Models with $\mu$-$\tau$ symmetry have been extensively studied \cite{Fukuyama:1997, Mohapatra:1999, Ma:2001a, Lam:2001, Kitabayashi:2003, Grimus:2003, Koide:2004, Ghosal:2003, Grimus:2005a, deGouvea:2004, Mohapatra:2005, Kitabayashi:2005, Mohapatra:2005a, Mohapatra:2005b, Mohapatra:2005c, Ahn:2006, Ge:2010}. The particularly important case of TB mixing is obtained when $\sin^2{2\theta_{12}}=8/9$ or $x+y=w+z$ \footnote{The other solution $x-y=w+z$ gives rise to TB mixing in another phase convention and is physically equivalent to $x+y=w+z$.} . In this case the matrix $m_\nu$ takes the form: \begin{equation} m_\nu=\left(\matrix{ x&y&y\cr y&x+v&y-v\cr y&y-v&x+v}\right)~~~, \label{gl21} \end{equation} In fact, in this case, $U=U_{TB}$ is given by \cite{Harrison:2002, Harrison:2002a, Harrison:2003, Harrison:2004}: \begin{equation} U_{TB}= \left(\matrix{ \displaystyle\sqrt{\frac{2}{3}}&\displaystyle\frac{1}{\sqrt 3}&0\cr -\displaystyle\frac{1}{\sqrt 6}&\displaystyle\frac{1}{\sqrt 3}&-\displaystyle\frac{1}{\sqrt 2}\cr -\displaystyle\frac{1}{\sqrt 6}&\displaystyle\frac{1}{\sqrt 3}&\displaystyle\frac{1}{\sqrt 2}}\right)~~~, \label{2} \end{equation} and, from eq. (\ref{numass}) one obtains: \begin{equation} m_{\nu}= m_1\Phi_1 \Phi_1^T + m_2\Phi_2 \Phi_2^T + m_3\Phi_3 \Phi_3^T~~~, \label{1k1} \end{equation} where \begin{equation} \Phi_1^T=\frac{1}{\sqrt{6}}(2,-1,-1)~~~,~~~~~ \Phi_2^T=\frac{1}{\sqrt{3}}(1,1,1)~~~,~~~~~\Phi_3^T=\frac{1}{\sqrt{2}}(0,-1,1) \label{4k1} \end{equation} are the respective columns of $U_{TB}$ and $m_i$ are the neutrino mass eigenvalues ($m_1=x-y$, $m_2=x+2y$ and $m_3=x-y+2v$). It is easy to see that the TB mass matrix in eqs. (\ref{1k1},\ref{4k1}) is indeed of the form in eq. (\ref{gl21}). All patterns for the neutrino spectrum are in principle possible. For a hierarchical spectrum $m_3>>m_2>>m_1$, $m_3^2 \sim \Delta m^2_{atm}$, $m_2^2/m_3^2 \sim \Delta m^2_{sol}/\Delta m^2_{atm}$ and $m_1$ could be negligible. But also degenerate masses and inverse hierarchy can be reproduced: for example, by taking $m_3= - m_2=m_1$ we have a degenerate model, while for $m_1= - m_2$ and $m_3=0$ an inverse hierarchy case is realized (stability under renormalization group running (for a review see for example \cite{Chankowski:2001mx}) strongly prefers opposite signs for the first and the second eigenvalue which are related to solar oscillations and have the smallest mass squared splitting). Note that the mass matrix for TB mixing, in the basis where charged leptons are diagonal, as given in eq. (\ref{gl21}), can be specified as the most general matrix which is invariant under $\mu-\tau$ (or 2-3) symmetry (see eqs. (\ref{mmutau}),(\ref{Amutau})) and, in addition, under the action of a unitary symmetric matrix $S_{TB}$ (actually $S_{TB}^2=1$ and $[S_{TB},A_{23}]=0$): \begin{eqnarray} m_\nu=S_{TB}m_\nu S_{TB}~~~,~~~~~m_\nu=A_{23}m_\nu A_{23}~~~, \label{inv} \end{eqnarray} where $S_{TB}$ is given by: \begin{eqnarray} \label{trep} S_{TB}&=\displaystyle\frac{1}{3} \left(\matrix{ -1&2&2\cr 2&-1&2\cr 2&2&-1}\right)~~~. \end{eqnarray} As a last example consider the case of BM where, in addition to $\theta_{13}=0$ and $\theta_{23}$ maximal, one also has $\sin^2{2\theta_{12}}=1$. The BM mixing matrix is given by: \begin{equation} U_{BM}= \left(\matrix{ \displaystyle\frac{1}{\sqrt 2}&\displaystyle-\frac{1}{\sqrt 2}&0\cr \displaystyle\frac{1}{2}&\displaystyle\frac{1}{2}&-\displaystyle\frac{1}{\sqrt 2}\cr \displaystyle\frac{1}{2}&\displaystyle\frac{1}{2}&\displaystyle\frac{1}{\sqrt 2}}\right)~~~. \label{21} \end{equation} In the basis where charged lepton masses are diagonal, from eq. (\ref{numass}), we derive the effective neutrino mass matrix in the BM case: \begin{equation} m_{\nu}= m_1\Phi_1 \Phi_1^T + m_2\Phi_2 \Phi_2^T + m_3\Phi_3 \Phi_3^T~~~, \label{1k} \end{equation} where \begin{equation} \Phi_1^T=\frac{1}{2}(\sqrt{2},1,1)~~~,~~~~~ \Phi_2^T=\frac{1}{2}(-\sqrt{2},1,1)~~~,~~~~~\Phi_3^T=\frac{1}{\sqrt{2}}(0,-1,1) \label{4k} \end{equation} As we see the most general mass matrix leading to BM mixing is of the form: \begin{equation} m_{\nu}=\left(\matrix{ x&y&y\cr y&z&x-z\cr y&x-z&z}\right)\;, \label{gl2} \end{equation} The resulting matrix can be completely characterized by the requirement of being invariant under the action of $A_{23}$ and also of the unitary, real, symmetric matrix $S_{BM}$ (satisfying $S_{BM}^2=1$ and $[S_{BM},A_{23}]=0$): \begin{equation} m_{\nu}= S_{BM} m_{\nu}S_{BM}~~~, ~~~~~~m_{\nu}=A_{23}m_{\nu}A_{23}~~~, \label{invS} \end{equation} with $S_{BM}$ given by: \begin{equation} S_{BM}=\left( \begin{array}{ccc} 0 & -\displaystyle\frac{1}{\sqrt{2}} & -\displaystyle\frac{1}{\sqrt{2}} \\ -\displaystyle\frac{1}{\sqrt{2}} & \displaystyle\frac{1}{2} & -\displaystyle\frac{1}{2} \\ -\displaystyle\frac{1}{\sqrt{2}} & -\displaystyle\frac{1}{2} & \displaystyle\frac{1}{2} \\ \end{array} \right)~~~. \label{matS} \end{equation} The $m_\nu$ mass matrices of the previous examples were all derived in the basis where charged leptons are diagonal. It is useful to consider the product $m^2=m_e^\dagger m_e$, where $m_e$ is the charged lepton mass matrix (defined as $\overline \psi_R m_e \psi_L$), because this product transforms as $m'^2=U_e^\dagger m^2 U_e$, with $U_e$ the unitary matrix that rotates the left-handed (LH) charged lepton fields. The most general diagonal $m^2$ is invariant under a diagonal phase matrix with 3 different phase factors: \begin{equation} m_e^\dagger m_e= T^\dagger m_e^\dagger m_e T \label{Tdiag} \end{equation} and conversely a matrix $m_e^\dagger m_e$ satisfying the above requirement is diagonal. If $T^n=1$ the matrix $T$ generates a cyclic group $Z_n$. In the simplest case n=3 and we get $Z_3$ but $n>3$ is equally possible. Examples are: \begin{eqnarray} \label{ta4} T_{TB}=\left(\matrix{ 1&0&0\cr 0&\omega&0\cr 0&0&\omega^2} \right). \end{eqnarray} where $\omega^3=1$, so that $T_{TB}^3=1$, or \begin{eqnarray} \label{ts4} T_{BM}=\left( \begin{array}{ccc} -1 & 0 & 0 \\ 0 & -i & 0 \\ 0 & 0 & i \\ \end{array} \right). \end{eqnarray} with $T_{BM}^4=1$. We are now in a position to explain the role of finite groups and to formulate the general strategy to obtain one of the previous special mass matrices, for example that of TB mixing. We must find a group $G_f$ which, for simplicity, must be as small as possible but large enough to contain the $S$ and $T$ transformations. A limited number of products of $S$ and $T$ close a finite group $G_f$. Hence the group $G_f$ contains the subgroups $G_S$ and $G_T$ generated by monomials in $S$ and $T$, respectively. We assume that the theory is invariant under the spontaneously broken symmetry described by $G_f$. Then we must arrange a breaking of $G_f$ such that, in leading order, $G_f$ is broken down to $G_S$ in the neutrino mass sector and down to $G_T$ in the charged lepton mass sector. In a good model this step must be realized in a natural way as a consequence of the stated basic principles, and not put in by hand. The symmetry under $A_{23}$ in some cases is also part of $G_f$ (this the case of $S_4$) and then must be preserved in the neutrino sector along with $S$ by the $G_f$ breaking or it could arise as a consequence of a special feature of the $G_f$ breaking (for example, in $A_4$ it is obtained by allowing only some transformation properties for the flavons with non vanishing VEV's). The explicit example of $A_4$ is discussed in the next section. Note that, along the same line, a model with $\mu-\tau$ symmetry can be realized in terms of the group $S_3$ generated by products of $A_{23}$ and $T$ (see, for example, \cite{Feruglio:2007a}). \section{The $A_4$ group} $A_4$ is the group of the even permutations of 4 objects. It has 4!/2=12 elements. Geometrically, it can be seen as the invariance group of a tetrahedron (the odd permutations, for example the exchange of two vertices, cannot be obtained by moving a rigid solid). Let us denote a generic permutation $(1,2,3,4)\rightarrow (n_1,n_2,n_3,n_4)$ simply by $(n_1n_2n_3n_4)$. $A_4$ can be generated by two basic permutations $S$ and $T$ given by $S=(4321)$ and $T=(2314)$. One checks immediately that: \begin{equation}\label{pres} S^2=T^3=(ST)^3=1 \end{equation} This is called a "presentation" of the group. The 12 even permutations belong to 4 equivalence classes ($h$ and $k$ belong to the same class if there is a $g$ in the group such that $ghg^{-1}=k$) and are generated from $S$ and $T$ as follows: \begin{eqnarray} \label{class} &C1&: I=(1234)\\ \nonumber &C2&: T=(2314),ST=(4132),TS=(3241),STS=(1423)\\ \nonumber &C3&: T^2=(3124),ST^2=(4213),T^2S=(2431),TST=(1342)\\ \nonumber &C4&: S=(4321),T^2ST=(3412),TST^2=(2143)\\ \nonumber \end{eqnarray} Note that, except for the identity $I$ which always forms an equivalence class in itself, the other classes are according to the powers of $T$ (in C4 $S$ could as well be seen as $ST^3$). The characters of a group $\chi_g^R$ are defined, for each element $g$, as the trace of the matrix that maps the element in a given representation $R$. From the invariance of traces under similarity transformations it follows that equivalent representations have the same characters and that characters have the same value for all elements in an equivalence class. Characters satisfy $\sum_g \chi_g^R \chi_g^{S*}= N \delta^{RS}$, where $N$ is the number of transformations in the group ($N=12$ in $A_4$). Also, for each element $h$, the character of $h$ in a direct product of representations is the product of the characters: $\chi_h^{R\otimes S}=\chi_h^R \chi_h^S$ and also is equal to the sum of the characters in each representation that appears in the decomposition of $R\otimes S$. In a finite group the squared dimensions of the inequivalent irreducible representations add up to $N$. The character table of $A_4$ is given in Table 3. From this table one derives that $A_4$ has four inequivalent representations: three of dimension one, $1$, $1'$ and $1"$ and one of dimension $3$. It is immediate to see that the one-dimensional unitary representations are obtained by: \begin{eqnarray} \label{uni} 1&S=1&T=1\\ \nonumber 1'&S=1&T=e^{\displaystyle i 2 \pi/3}\equiv\omega\\\nonumber 1''&S=1&T=e^{\displaystyle i 4\pi/3}\equiv\omega^2\\\nonumber \end{eqnarray} Note that $\omega=-1/2+i \sqrt{3}/2$ is the cubic root of 1 and satisfies $\omega^2=\omega^*$, $1+\omega+\omega^2=0$. \begin{table}[t] \begin{center} \begin{tabular}{|l|c|c|c|c|} \hline \textbf{Class} & \textbf{$\chi^1$} & \textbf{$\chi^{1'}$} &\textbf{$\chi^{1"}$}&\textbf{$\chi^3$}\\ \hline $C_1$ & 1 & 1& 1&3\\ \hline $C_2$ & 1 &$\omega$&$\omega^2$&0\\ \hline $C_3$ & 1 & $\omega^2$&$\omega$&0\\ \hline $C_4$ & 1 & 1& 1&-1\\ \hline \end{tabular} \caption{Characters of $A_4$} \label{tcar} \end{center} \end{table} The three-dimensional unitary representation, in a basis where the element $S=S'$ is diagonal, is built up from: \begin{equation}\label{tre} S'=\left(\matrix{ 1&0&0\cr 0&-1&0\cr 0&0&-1}\right),~ T'=\left(\matrix{ 0&1&0\cr 0&0&1\cr 1&0&0} \right). \end{equation} The multiplication rules are as follows: the product of two 3 gives $3 \times 3 = 1 + 1' + 1'' + 3 + 3$ and $1' \times 1' = 1''$, $1' \times 1'' = 1$, $1'' \times 1'' = 1'$ etc. If $3\sim (a_1,a_2,a_3)$ is a triplet transforming by the matrices in eq. (\ref{tre}) we have that under $S'$: $S'(a_1,a_2,a_3)^t= (a_1,-a_2,-a_3)^t$ (here the upper index $t$ indicates transposition) and under $T'$: $T'(a_1,a_2,a_3)^t= (a_2,a_3,a_1)^t$. Then, from two such triplets $3_a\sim (a_1,a_2,a_3)$, $3_b\sim (b_1,b_2,b_3)$ the irreducible representations obtained from their product are: \begin{equation} 1=a_1b_1+a_2b_2+a_3b_3 \end{equation} \begin{equation} 1'=a_1b_1+\omega^2 a_2b_2+\omega a_3b_3 \end{equation} \begin{equation} 1"=a_1b_1+\omega a_2b_2+\omega^2 a_3b_3 \end{equation} \begin{equation} 3\sim (a_2b_3, a_3b_1, a_1b_2) \end{equation} \begin{equation} 3\sim (a_3b_2, a_1b_3, a_2b_1) \end{equation} In fact, take for example the expression for $1"=a_1b_1+\omega a_2b_2+\omega^2 a_3b_3$. Under $S'$ it is invariant and under $T'$ it goes into $a_2b_2+\omega a_3b_3+\omega^2 a_1b_1=\omega^2[a_1b_1+\omega a_2b_2+\omega^2 a_3b_3]$ which is exactly the transformation corresponding to $1"$. In eq. (\ref{tre}) we have the representation 3 in a basis where $S$ is diagonal. We shall see that for our purposes it is convenient to go to a basis where instead it is $T$ that is diagonal. This is obtained through the unitary transformation: \begin{eqnarray} T&=&VT'V^\dagger=\left(\matrix{ 1&0&0\cr 0&\omega&0\cr 0&0&\omega^2} \right),\label{mT}\\ S&=&VS'V^\dagger=\frac{1}{3} \left(\matrix{ -1&2&2\cr 2&-1&2\cr 2&2&-1}\right). \label{mS} \end{eqnarray} where: \begin{equation} \label {vu} V=\frac{1}{\sqrt{3}} \left(\matrix{ 1&1&1\cr 1&\omega^2&\omega\cr 1&\omega&\omega^2}\right). \end{equation} The matrix $V$ is special in that it is a 3x3 unitary matrix with all entries of unit absolute value. It is interesting that this matrix was proposed long ago as a possible mixing matrix for neutrinos \cite{Cabibbo:1978, Wolfenstein:1978}. We shall see in the following that in the $T$ diagonal basis the charged lepton mass matrix (to be precise the matrix $m_e^\dagger m_e$) is diagonal. Notice that the matrices $(S,T)$ of eqs. (\ref{mT}-\ref{mS}) coincide with the matrices $(S_{TB},T_{TB})$ of the previous section. In this basis the product rules of two triplets, ($\psi_1,\psi_2,\psi_3$) and ($\varphi_1,\varphi_2,\varphi_3$) of $A_4$, according to the multiplication rule $3\times 3=1+1'+1"+3+3$ are different than in the $S$ diagonal basis (because for Majorana mass matrices the relevant scalar product is $(ab)$ and not $(a^\dagger b$) and are given by: \begin{eqnarray} \label{tensorproda4} &\psi_1\varphi_1+\psi_2\varphi_3+\psi_3\varphi_2 \sim 1 ~,\nonumber \\ &\psi_3\varphi_3+\psi_1\varphi_2+\psi_2\varphi_1 \sim 1' ~,\nonumber \\ &\psi_2\varphi_2+\psi_3\varphi_1+\psi_1\varphi_3 \sim 1'' ~,\nonumber \end{eqnarray} \begin{equation} \left( \begin{array}{c} 2\psi_1\varphi_1-\psi_2\varphi_3-\psi_3\varphi_2 \\ 2\psi_3\varphi_3-\psi_1\varphi_2-\psi_2\varphi_1 \\ 2\psi_2\varphi_2-\psi_1\varphi_3-\psi_3\varphi_1 \\ \end{array} \right) \sim 3_S~, \qquad \left( \begin{array}{c} \psi_2\varphi_3-\psi_3\varphi_2 \\ \psi_1\varphi_2-\psi_2\varphi_1 \\ \psi_3\varphi_1-\psi_1\varphi_3 \\ \end{array} \right) \sim 3_A~. \label{tensorp} \end{equation} In the following we will work in the $T$ diagonal basis, unless otherwise stated. In this basis the 12 matrices of the 3-dimensional representation of $A_4$ are given by: \begin{center} \begin{tabular}{ll} ${\cal C}} \def\cD{{\cal D}_1$ : &$1= \left( \begin{array}{ccc} 1 & 0 & 0 \\ 0 & 1 & 0 \\ 0 & 0 & 1 \\ \end{array} \right)$, \\ \\ ${\cal C}} \def\cD{{\cal D}_2$ : &$T= \left( \begin{array}{ccc} 1 & 0 & 0 \\ 0 & \omega & 0 \\ 0 & 0 & \omega^2 \\ \end{array} \right), ST=\frac{1}{3}\left( \begin{array}{ccc} -1 & 2\omega & 2\omega^2 \\ 2 & -\omega & 2\omega^2 \\ 2 & 2\omega & -\omega^2 \\ \end{array} \right), $,\\[0.6cm] & $TS=\frac{1}{3}\left( \begin{array}{ccc} -1 & 2 & 2 \\ 2\omega & -\omega & 2\omega \\ 2\omega^2 & 2\omega^2 & -\omega^2 \\ \end{array} \right), STS=\frac{1}{3}\left( \begin{array}{ccc} -1 & 2\omega^2 & 2\omega \\ 2\omega^2 & -\omega & 2 \\ 2\omega & 2 & -\omega^2 \\ \end{array} \right)$, \\ \\ ${\cal C}} \def\cD{{\cal D}_3$ : & $T^2=\left( \begin{array}{ccc} 1 & 0 & 0 \\ 0 & \omega^2 & 0 \\ 0 & 0 & \omega \\ \end{array} \right), ST^2=\frac{1}{3}\left( \begin{array}{ccc} -1 & 2\omega^2 & 2\omega \\ 2 & -\omega^2 & 2\omega \\ 2 & 2\omega^2 & -\omega \\ \end{array} \right)$,\\[0.6cm] &$T^2S=\frac{1}{3}\left( \begin{array}{ccc} -1 & 2 & 2 \\ 2\omega^2 & -\omega^2 & 2\omega^2 \\ 2\omega & 2\omega & -\omega \\ \end{array} \right), TST=\frac{1}{3}\left( \begin{array}{ccc} -1 & 2\omega & 2\omega^2 \\ 2\omega & -\omega^2 & 2 \\ 2\omega^2 & 2 & -\omega \\ \end{array} \right)$, \end{tabular} \end{center} \begin{center} \begin{tabular}{ll} ${\cal C}} \def\cD{{\cal D}_4$ : & $S=\frac{1}{3}\left( \begin{array}{ccc} -1 & 2 & 2 \\ 2 & -1 & 2 \\ 2 & 2 & -1 \\ \end{array} \right), T^2ST=\frac{1}{3}\left( \begin{array}{ccc} -1 & 2\omega & 2\omega^2 \\ 2\omega^2 & -1 & 2\omega \\ 2\omega & 2\omega^2 & -1 \\ \end{array} \right)$,\\[0.6cm] &$ TST^2=\frac{1}{3}\left( \begin{array}{ccc} -1 & 2\omega^2 & 2\omega \\ 2\omega & -1 & 2\omega^2 \\ 2\omega^2 & 2\omega & -1 \\ \end{array} \right)$. \end{tabular} \end{center} We can now see why $A_4$ works for TB mixing. In section 2 we have already mentioned that the most general mass matrix for TB mixing in eq. (\ref{gl21}), in the basis where charged leptons are diagonal, can be specified as one which is invariant under the 2-3 (or $\mu-\tau$) symmetry and under the $S$ unitary transformation, as stated in eq. (\ref{inv}) (note that $S_{TB}$ in eqs. (\ref{inv}, \ref{trep}) coincides with $S$ in eq. (\ref{mS})). This observation plays a key role in leading to $A_4$ as a candidate group for TB mixing, because $S$ is a matrix of $A_4$. Instead the matrix $A_{23}$ is not an element of $A_4$ (because the 2-3 exchange is an odd permutation). We shall see that in $A_4$ models the 2-3 symmetry is maintained by imposing that there are no flavons transforming as $1'$ or $1''$ that break $A_4$ with two different VEV's (in particular one can assume that there are no flavons in the model transforming as $1'$ or $1''$). It is also clear that a generic diagonal charged lepton matrix $m_e^\dagger m_e$ is characterized by the invariance under $T$, or $T^\dagger m_e^\dagger m_e T=m_e^\dagger m_e$. The group $A_4$ has two obvious subgroups: $G_S$, which is a reflection subgroup generated by $S$ and $G_T$, which is the group generated by $T$, which is isomorphic to $Z_3$. If the flavor symmetry associated to $A_4$ is broken by the VEV of a triplet $\varphi=(\varphi_1,\varphi_2,\varphi_3)$ of scalar fields, there are two interesting breaking pattern. The VEV \begin{equation} \langle\varphi\rangle=(v_S,v_S,v_S) \label{unotre} \end{equation} breaks $A_4$ down to $G_S$, while \begin{equation} \langle\varphi\rangle=(v_T,0,0) \label{unozero} \end{equation} breaks $A_4$ down to $G_T$. As we will see, $G_S$ and $G_T$ are the relevant low-energy symmetries of the neutrino and the charged-lepton sectors, respectively. Indeed we have already seen that the TB mass matrix is invariant under $G_S$ and a diagonal charged lepton mass $m_e^\dagger m_e$ is invariant under $G_T$. \section{Applying $A_4$ to lepton masses and mixings} In the lepton sector a typical $A_4$ model works as follows \cite{Altarelli:2006}. One assigns leptons to the four inequivalent representations of $A_4$: LH lepton doublets $l$ transform as a triplet $3$, while the RH charged leptons $e^c$, $\mu^c$ and $\tau^c$ transform as $1$, $1''$ and $1'$, respectively. Here we consider a see-saw realization, so we also introduce conjugate neutrino fields $\nu^c$ transforming as a triplet of $A_4$. We adopt a supersymmetric (SUSY) context also to make contact with Grand Unification (flavor symmetries are supposed to act near the GUT scale). In fact, as well known, SUSY is important in GUT's for offering a solution to the hierarchy problem, for improving coupling unification and for making the theory compatible with bounds on proton decay. But in models of lepton mixing SUSY also helps for obtaining the vacuum alignment, because the SUSY constraints are very strong and limit the form of the superpotential very much. Thus SUSY is not necessary but it is a plausible and useful ingredient. The flavor symmetry is broken by two triplets $\varphi_S$ and $\varphi_T$ and by one or more singlets $\xi$. All these fields are invariant under the SM gauge symmetry. Two Higgs doublets $h_{u,d}$, invariant under $A_4$, are also introduced. One can obtain the observed hierarchy among $m_e$, $m_\mu$ and $m_\tau$ by introducing an additional U(1)$_{FN}$ flavor symmetry \cite{Froggatt:1979} under which only the RH lepton sector is charged (recently some models were proposed with a different VEV alignment such that the charged lepton hierarchies are obtained without introducing a $U(1)$ symmetry \cite{Lin:2009, Lin:2009a, Altarelli:2009a}). We recall that $U(1)_{FN}$ is a simplest flavor symmetry where particles in different generations are assigned (in general) different values of an Abelian charge. Also Higgs fields may get a non zero charge. When the symmetry is spontaneously broken the entries of mass matrices are suppressed if there is a charge mismatch and more so if the corresponding mismatch is larger. We assign FN-charges $0$, $q$ and $2q$ to $\tau^c$, $\mu^c$ and $e^c$, respectively. There is some freedom in the choice of $q$. Here we take $q=2$. By assuming that a flavon $\theta$, carrying a negative unit of FN charge, acquires a VEV $\langle \theta \rangle/\Lambda\equiv\lambda<1$, the Yukawa couplings become field dependent quantities $y_{e,\mu,\tau}=y_{e,\mu,\tau}(\theta)$ and we have \begin{equation} y_\tau\approx \mathcal{O}(1)~~~,~~~~~~~y_\mu\approx O(\lambda^2)~~~, ~~~~~~~y_e\approx O(\lambda^{4})~~~. \end{equation} Had we chosen $q=1$, we would have needed $\langle \theta \rangle/\Lambda$ of order $\lambda^2$, to reproduce the above result. The superpotential term for lepton masses, $w_l$ is given by: \begin{equation} w_l=y_e e^c (\varphi_T l)+y_\mu \mu^c (\varphi_T l)'+ y_\tau \tau^c (\varphi_T l)''+ y (\nu^c l)+ (x_A\xi+\tilde{x}_A\tilde{\xi}) (\nu^c\nu^c)+x_B (\varphi_S \nu^c\nu^c)+h.c.+... \label{wlss} \end{equation} with dots denoting higher dimensional operators that lead to corrections to the LO approximation. In our notation, the product of 2 triplets $(3 3)$ transforms as $1$, $(3 3)'$ transforms as $1'$ and $(3 3)''$ transforms as $1''$. To keep our formulae compact, we omit to write the Higgs and flavon fields $h_{u,d}$, $\theta$ and the cut-off scale $\Lambda$. For instance $y_e e^c (\varphi_T l)$ stands for $y_e e^c (\varphi_T l) h_d \theta^4/\Lambda^5$. The parameters of the superpotential $w_l$ are complex, in particular those responsible for the heavy neutrino Majorana masses, $x_{A,B}$. Some terms allowed by the $A_4$ symmetry, such as the terms obtained by the exchange $\varphi_T\leftrightarrow \varphi_S$, (or the term $(\nu^c\nu^c)$) are missing in $w_l$. Their absence is crucial and, in each version of $A_4$ models, is motivated by additional symmetries (In ref. \cite{Altarelli:2005} a natural solution of this problem based on a formulation with extra dimensions was discussed; for a similar approach see also \cite{Csaki:2008,Kadosh:2010rm}). In the present version the additional symmetry is $Z_3$. A $U(1)_R$ symmetry related to R-parity and the presence of driving fields in the flavon superpotential are common features of supersymmetric formulations. Eventually, after the inclusion of N = 1 SUSY breaking effects, the $U(1)_R$ symmetry will be broken at the low energy scale $m_{SUSY}$ down to the discrete R-parity. Supersymmetry also helps producing and maintaining the hierarchy $\langle h_{u,d}\rangle=v_{u,d}\ll \Lambda$ where $\Lambda$ is the cut-off scale of the theory. The fields in the model and their classification under the symmetry are summarized in Table \ref{table:TransformationsA}. \begin{table}[h] \begin{center} \begin{tabular}{|c||c|c|c|c|c|c||c|c|c|c||c|c|c|} \hline &&&&&&&&&&&&&\\[-0,3cm] & $l$ & $e^c$ & $\mu^c$ & $\tau^c$ & $\nu^c$ & $h_{u,d}$ & $\theta$ & $\varphi_T$ & $\varphi_S$ & $\xi$ &$\varphi_0^T$ & $\varphi_0^S$ & $\xi_0$ \\ &&&&&&&&&&&&&\\[-0,3cm] \hline &&&&&&&&&&&&&\\[-0,3cm] $A_4$ & 3 & 1 & $1''$ & $1'$ & 3 & 1 & 1 & 3 & $3$ & 1 & 3 & 3 & 1 \\ &&&&&&&&&&&&&\\[-0,3cm] $Z_3$ & $\omega$ &$\omega^2$ & $\omega^2$& $\omega^2$& $\omega^2$ & 1 & 1 &1& $\omega^2$&$\omega^2$& 1 &$\omega^2$ & $\omega^2$ \\ &&&&&&&&&&&&&\\[-0,3cm] $U(1)_{FN}$ & 0 & 4 & 2 & 0 & 0 & 0 & -1 & 0 & 0 & 0 & 0 & 0 & 0 \\ &&&&&&&&&&&&&\\[-0,3cm] $U(1)_R$ & 1 & 1 & 1 & 1 & 1 & 0 & 0 & 0 & 0 & 0 & 2 & 2 & 2 \\ \hline \end{tabular} \end{center} \caption{\label{table:TransformationsA}Transformation properties of all the fields.} \end{table} In this set up it can be shown that the fields $\varphi_T$, $\varphi_S$ and $\xi$ develop a VEV along the directions: \begin{eqnarray} \langle \varphi_T \rangle&=&(v_T,0,0)\nonumber\\ \langle \varphi_S\rangle&=&(v_S,v_S,v_S)\nonumber\\ \langle \xi \rangle&=&u~~~. \label{align} \end{eqnarray} A crucial part of all serious $A_4$ models is the dynamical generation of this alignment in a natural way. We refer to ref. \cite{Altarelli:2006} for a proof that the above alignment naturally follows from the most general LO superpotential implied by the symmetries of the model. As already mentioned, the group $A_4$ has two obvious subgroups: $G_S$, which is a reflection subgroup generated by $S$ and $G_T$, which is the group generated by $T$, isomorphic to $Z_3$. In the basis where $S$ and $T$ are given by eq. (\ref{trep}), the VEV $\langle \varphi_T \rangle=(v_T,0,0)$ breaks $A_4$ down to $G_T$, while $\langle \varphi_S\rangle=(v_S,v_S,v_S)$ breaks $A_4$ down to $G_S$. If the alignment in eq. (\ref{align}) is realized, at the leading order of the $1/\Lambda$ expansion, the mass matrices $m_l$ and $m_\nu$ for charged leptons and neutrinos correspond to TB mixing. The charged lepton mass matrix is diagonal: \begin{equation} m_l=v_d\frac{v_T}{\Lambda}\left( \begin{array}{ccc} y_e& 0& 0\\ 0& y_\mu & 0 \\ 0& 0& y_\tau \end{array} \right)~~~, \label{mch} \end{equation} The charged fermion masses are given by: \begin{equation} \label{chmasses} m_e= y_e v_d \frac{v_T}{\Lambda}~~~,~~~~~~~ m_\mu= y_\mu v_d \frac{v_T}{\Lambda}~~~,~~~~~~~ m_\tau=y_\tau v_d \frac{v_T}{\Lambda}~~~, \end{equation} where the suppression coming from the breaking of $U(1)_{FN}$ is understood. For example $y_e$ stands for $y_e \theta^4/\Lambda^4$. In the neutrino sector, after electroweak and $A_4$ symmetry breaking we have Dirac and Majorana masses: \begin{equation} m_\nu^D=\left( \begin{array}{ccc} 1 & 0 & 0 \\ 0 & 0 & 1 \\ 0 & 1 & 0 \\ \end{array} \right)yv_u\qquad\qquad,~~ M=\left( \begin{array}{ccc} A+2 B/3& -B/3& -B/3\\ -B/3& 2B/3& A-B/3\\ -B/3& A-B/3& 2 B/3 \end{array} \right) u ~~~, \end{equation} where \begin{equation} A\equiv 2 x_A ~~~,~~~~~~~B\equiv 2 x_B \frac{v_S}{u}~~~. \label{add} \end{equation} The eigenvalues of $M$ are \begin{equation} M_1=(A+B)u~~~,~~~~~~~ M_2=Au~~~,~~~~~~~M_3=(-A+B)u. \end{equation} The mass matrix for light neutrinos is $m_\nu=(m^D_\nu)^T M^{-1} m^D_\nu$ with eigenvalues \begin{equation} m_1=\frac{y^2 v_u^2}{M_1}~~~,~~~~~~~ m_2=\frac{y^2 v_u^2}{M_2}~~~,~~~~~~~ m_3=\frac{y^2 v_u^2}{M_3}~~~. \label{nueig} \end{equation} The mixing matrix is $U_{TB}$, eq. (\ref{2}). Both normal and inverted hierarchies in the neutrino mass spectrum can be realized. It is interesting that $A_4$ models with the see-saw mechanism typically lead to a light neutrino spectrum which satisfies the sum rule (among complex masses): \begin{equation} \frac{1}{m_3}=\frac{1}{m_1}-\frac{2}{m_2}~~~.\\ \label{sumr} \end{equation} The phases of the complex parameters $A$ and $B$ do not produce any CP violation in neutrino oscillations, since $\theta_{13}=0$, but are quite important to make the above sum rule compatible with the present data on neutrino masses. A detailed discussion of a spectrum of this type can be found in refs. \cite{Altarelli:2006,Altarelli:2009a}. Both types of ordering, normal and inverted are allowed and the above sum rule gives rise to bounds on the lightest neutrino mass. For normal ordering we have \begin{eqnarray} m_1&\ge& \sqrt{\frac{\Delta m^2_{sun}}{3}} \left(1-\displaystyle\frac{4\sqrt{3}}{9} r+ ...\right) \approx 0.004~{\rm eV}\nonumber\\ m_1&\le& \sqrt{\frac{\Delta m^2_{sun}}{3}}\left(1+\displaystyle\frac{4\sqrt{3}}{9} r+...\right) \approx 0.006~{\rm eV} \label{boundno} \end{eqnarray} and for the inverted ordering: \begin{equation} m_3\ge \sqrt{\frac{\Delta m^2_{atm}}{8}} \left(1-\displaystyle\frac{1}{6} r^2+...\right) \approx 0.017~{\rm eV}, \label{boundio} \end{equation} where the dots represent terms with higher powers of $r$. Notice that for normal ordering the neutrino mass spectrum is essentially determined: $m_1\approx 0.005$ eV, $m_1\approx 0.01$ eV and $m_3\approx 0.05$ eV. Also the possible values of $|m_{ee}|$ are restricted. For normal hierarchy we have \begin{equation} |m_{ee}|\approx \displaystyle\frac{4}{3\sqrt{3}} \Delta m^2_{sun} \approx 0.007~{\rm eV}~~~. \label{meeno} \end{equation} while for inverted hierarchy \begin{equation} |m_{ee}|\ge \displaystyle\sqrt{\frac{\Delta m^2_{atm}}{8}} \approx 0.017~{\rm eV}~~~. \label{meeio} \end{equation} In a completely general framework, without the restrictions imposed by the flavor symmetry, $|m_{ee}|$ could vanish in the case of normal hierarchy. In this model $|m_{ee}|$ is always different from zero, though its value for normal hierarchy is probably too small to be detected in the next generation of $0\nu\beta\beta$ experiments. Note that in the charged lepton sector the flavor symmetry $A_4$ is broken by $\langle \varphi_T \rangle$ down to $G_T$. Actually the above mass terms for charged leptons are the most general allowed by the symmetry $G_T$. At leading order in $1/\Lambda$, charged lepton masses are diagonal simply because there is a low-energy $G_T$ symmetry. In the neutrino sector $A_4$ is broken down to $G_S$, though neutrino masses in this model are not the most general ones allowed by $G_S$. The additional property which is needed, the invariance under $A_{\mu\tau}$, is obtained by stipulating that there are no $A_4$ breaking flavons transforming like $1'$ and $1''$. In fact, from eq. (\ref{tensorproda4}), we see that the expressions for $(33)'$ and $(33)''$ are not 2-3 symmetric. At the next level of approximation each term of the superpotential is corrected by operators of higher dimension whose contributions are suppressed by at least one power of VEV's/$\Lambda$. The corrections to the relevant part of the superpotential determine small deviations from the LO VEV alignment configuration. The NLO corrections to mass and mixing matrices are obtained by inserting the corrected VEV alignment in the LO operators plus the contribution of the new operators evaluated with the unperturbed VEV's. The final result is \cite{Altarelli:2006} that, when the NLO corrections are included, TB mixing is violated by small terms of the same order for all mixing angles: \begin{eqnarray} \sin^2\theta_{12}&=&\frac{1}{3}+{\cal O}(\varepsilon)\nonumber \\ \sin^2\theta_{23}&=&\frac{1}{2}+{\cal O}(\varepsilon) \label{corr}\\ \sin\theta_{13}&=&{\cal O}(\varepsilon)\nonumber \end{eqnarray} where $\varepsilon$ is of order of the typical VEV in units of $\Lambda$. The fact that TB mixing is well satisfied by the data sets the restriction $\varepsilon < {\cal O}(\lambda_C^2)$. From the requirement that the Yukawa coupling $y_\tau$ remains in the perturbative regime, we also get a lower bound on $\varepsilon$ of about $0.01$, the exact value depending on $\tan\beta=v_u/v_d$ and on the largest allowed $|y_\tau|$. Thus we approximately have \begin{equation} 0.01<\varepsilon<0.05~~~. \label{range} \end{equation} From the see-saw relations in eq. (\ref{nueig}), assuming a coupling $y$ of order one, we see that the heavy RH neutrino masses are all of order $10^{15}$ GeV, close to the GUT scale. The cut-off of the theory can be estimated form eq. (\ref{range}) to be close to $10^{17}$ GeV. The above results in eqs. (\ref{mch}-\ref{meeio}) on the lepton mass matrices and the neutrino spectrum refer to the LO approximation. Relations among neutrino masses can be affected by NLO corrections but, for $\varepsilon$ varying in the range of eq. (\ref{range}), the bounds (\ref{boundno},\ref{boundio}) do not appreciably change (see for example ref. \cite{Barry:2010zk} for a numerical study of the deviations induced by vacuum misalignment). Also corrections induced by the renormalization group evolution of the parameters can modify the above predictions, but only in the case of sufficiently degenerate mass levels $m_1$ and $m_2$ with equal phases, which occurs for inverted mass ordering and far from the lower bound (\ref{boundio}) \cite{Lin:2009b}. The expansion parameter $\varepsilon$ directly controls also other observables, such as the CP asymmetries of leptogenesis and the rates of lepton flavor violating transitions. This provides an interesting link between the physics in the early universe relevant for leptogenesis and the low energy physics accessible in current experiments. We will discuss the interplay bewteen discrete flavor symmetries and leptogenesis in Sect. 10. \section{Possible origin of $A_4$} There is an interesting relation \cite{Altarelli:2006} between the $A_4$ model considered so far and the modular group. This relation could possibly be relevant to understand the origin of the $A_4$ symmetry from a more fundamental layer of the theory. The modular group $\Gamma$ is the group of linear fractional transformations acting on a complex variable $z$: \begin{equation} z\to\frac{az+b}{cz+d}~~~,~~~~~~~ad-bc=1~~~, \label{frac} \end{equation} where $a,b,c,d$ are integers. There are infinite elements in $\Gamma$, but all of them can be generated by the two transformations: \begin{equation} s:~~~z\to -\frac{1}{z}~~~,~~~~~~~t:~~~z\to z+1~~~, \label{st} \end{equation} The transformations $s$ and $t$ in (\ref{st}) satisfy the relations \begin{equation} s^2=(st)^3=1 \label{absdef} \end{equation} and, conversely, these relations provide an abstract characterization of the modular group. Since the relations (\ref{pres}) are a particular case of the more general constraint (\ref{absdef}), it is clear that $A_4$ is a very small subgroup of the modular group and that the $A_4$ representations discussed above are also representations of the modular group. In string theory the transformations (\ref{st}) operate in many different contexts. For instance the role of the complex variable $z$ can be played by a field, whose VEV can be related to a physical quantity like a compactification radius or a coupling constant. In that case $s$ in eq. (\ref{st}) represents a duality transformation and $t$ in eq. (\ref{st}) represent the transformation associated to an ''axionic'' symmetry. A different way to understand the dynamical origin of $A_4$ was recently presented in ref. \cite{Altarelli:2007} where it is shown that the $A_4$ symmetry can be simply obtained by orbifolding starting from a model in 6 dimensions (6D). In this approach $A_4$ appears as the remnant of the reduction from 6D to 4D space-time symmetry induced by the special orbifolding adopted. This approach suggests a deep relation between flavor symmetry in 4D and space-time symmetry in extra dimensions. The orbifolding is defined as follows. We consider a quantum field theory in 6 dimensions, with two extra dimensions compactified on an orbifold $T^2/Z_2$. We denote by $z=x_5+i x_6$ the complex coordinate describing the extra space. The torus $T^2$ is defined by identifying in the complex plane the points related by \begin{equation} \begin{array}{l} z\to z+1\\ z\to z+\gamma~~~~~~~~~~~~~~~~~\gamma=e^{\displaystyle i\frac{\pi}{3}}~~~, \label{torus} \end{array} \end{equation} where our length unit, $2\pi R$, has been set to 1 for the time being. The parity $Z_2$ is defined by \begin{equation} z\to -z \label{parity} \end{equation} and the orbifold $T^2/Z_2$ can be represented by the fundamental region given by the triangle with vertices $0,1,\gamma$, see Fig. 3. The orbifold has four fixed points, $(z_1,z_2,z_3,z_4)=(1/2,(1+\gamma)/2,\gamma/2,0)$. The fixed point $z_4$ is also represented by the vertices $1$ and $\gamma$. In the orbifold, the segments labelled by $a$ in Fig. 1, $(0,1/2)$ and $(1,1/2)$, are identified and similarly for those labelled by $b$, $(1,(1+\gamma)/2)$ and $(\gamma,(1+\gamma)/2)$, and those labelled by $c$, $(0,\gamma/2)$, $(\gamma,\gamma/2)$. Therefore the orbifold is a regular tetrahedron with vertices at the four fixed points. \begin{figure}[] \centering $$\hspace{-4mm} \includegraphics[width=10.0 cm]{tetrahedron.eps}$$ \caption[]{Orbifold $T_2/Z_2$. The regions with the same numbers are identified with each other. The four triangles bounded by solid lines form the fundamental region, where also the edges with the same letters are identified. The orbifold $T_2/Z_2$ is exactly a regular tetrahedron with 6 edges $a,b,c,d,e,f$ and four vertices $z_1$, $z_2$, $z_3$, $z_4$, corresponding to the four fixed points of the orbifold. } \end{figure} The symmetry of the uncompactified 6D space time is broken by compactification. Here we assume that, before compactification, the space-time symmetry coincides with the product of 6D translations and 6D proper Lorentz transformations. The compactification breaks part of this symmetry. However, due to the special geometry of our orbifold, a discrete subgroup of rotations and translations in the extra space is left unbroken. This group can be generated by two transformations: \begin{equation} \begin{array}{ll} {\cal S}:& z\to z+\frac{1}{2}\\ {\cal T}:& z\to \omega z~~~~~~~~~~~~~\omega\equiv\gamma^2~~~~. \label{rototra} \end{array} \end{equation} Indeed ${\cal S}$ and ${\cal T}$ induce even permutations of the four fixed points: \begin{equation} \begin{array}{cc} {\cal S}:& (z_1,z_2,z_3,z_4)\to (z_4,z_3,z_2,z_1)\\ {\cal T}:& (z_1,z_2,z_3,z_4)\to (z_2,z_3,z_1,z_4) \end{array}~~~, \label{stfix} \end{equation} thus generating the group $A_4$. From the previous equations we immediately verify that ${\cal S}$ and ${\cal T}$ satisfy the characteristic relations obeyed by the generators of $A_4$: ${\cal S}^2={\cal T}^3=({\cal ST})^3=1$. These relations are actually satisfied not only at the fixed points, but on the whole orbifold, as can be easily checked from the general definitions of ${\cal S}$ and ${\cal T}$ in eq. (\ref{rototra}), with the help of the orbifold defining rules in eqs. (\ref{torus}) and (\ref{parity}). We can exploit this particular geometry of the internal space to build a model with $A_4$ flavor symmetry. There are 4D branes at the four fixed points of the orbifolding and the tetrahedral symmetry of $A_4$ connects these branes. The standard model fields have components on the fixed point branes while the scalar fields necessary for the $A_4$ breaking are in the bulk. Each brane field, either a triplet or a singlet, has components on all of the four fixed points (in particular all components are equal for a singlet) but the interactions are local, i.e. all vertices involve products of field components at the same space-time point. In the low-energy limit this model coincides with the one illustrated in the previous section. Unfortunately in such a limit the 6D construction does not provide additional constraints or predictions. This construction can be embedded in a SU(5) GUT \cite{Burrows:2009}. Other discrete groups can arise from the compactification of two extra dimensions on orbifolds and the possibilities have been classified in \cite{Adulpravitchai:2009b,Adulpravitchai:2010na} within a field theory approach. In string theory the flavor symmetry can be larger than the isometry of the compact space. For instance in heterotic orbifold models the orbifold geometry combines with the space group selection rules of the string, as shown in \cite{Kobayashi:2007}. Discrete flavor symmetries from magnetized/intersecting D-branes are discussed in \cite{Abe:2009}. Discrete symmetries can also arise from the spontaneous breaking of continuous ones. Such a possibility has been discussed in ref. \cite{Adulpravitchai:2009c,Berger:2009}. \section{Alternative routes to TB mixing} While $A_4$ is the minimal flavor group leading to TB mixing, alternative flavor groups have been studied in the literature and can lead to interesting variants with some specific features. Recently, in ref. \cite{Lam:2008}, the claim was made that, in order to obtain the TB mixing "without fine tuning", the finite group must be $S_4$ or a larger group containing $S_4$. For us this claim is not well grounded being based on an abstract mathematical criterium for a natural model (see also \cite{Grimus:2008a}). For us a physical field theory model is natural if the interesting results are obtained from the most general lagrangian compatible with the stated symmetry and the specified representation content for the flavons. For example, we obtain from $A_4$ (which is a subgroup of $S_4$) a natural (in our sense) model for the TB mixing by simply not including symmetry breaking flavons transforming like the $1'$ and the $1''$ representations of $A_4$. This limitation on the transformation properties of the flavons is not allowed by the rules specified in ref. \cite{Lam:2008} which demand that the symmetry breaking is induced by all possible kinds of flavons (note that, according to this criterium, the SM of electroweak interactions would not be natural because only Higgs doublets are introduced!). Rather, for naturalness we also require that additional physical properties like the VEV alignment or the hierarchy of charged lepton masses also follow from the assumed symmetry and are not obtained by fine tuning parameters: for this actually $A_4$ can be more effective than $S_4$ because it possesses three different singlet representations 1, $1'$ and $1''$. Models of neutrino mixing based on $S_4$ have in fact been studied \cite{Mohapatra:2004, Hagedorn:2006, Cai:2006, Ma:2007a, Bazzocchi:2008c, Ishimori:2009, Bazzocchi:2009a, Bazzocchi:2009b, Meloni:2009, Dutta:2009, Dutta:2009a, Morisi:2010, Ding:2010,Hagedorn:2010th,Ishimori:2010xk}. The group of the permutations of 4 objects $S_4$ has 24 elements and 5 equivalence classes (the character table is given in Table \ref{tcars}) that correspond to 5 inequivalent irreducible representations, two singlets, one doublet, two triplets: $1_1$, $1_2$, $2$, $3_1$ and $3_2$ (see Table 2). Note that the squares of the dimensions af all these representations add up to 24. \begin{table}[h] \begin{center} \begin{tabular}{|l|c|c|c|c|c|} \hline \textbf{Class} & \textbf{$\chi(1_1)$} & \textbf{$\chi(1_2)$} &\textbf{$\chi(2)$}&\textbf{$\chi(3_1)$}&\textbf{$\chi(3_2)$}\\ \hline $C_1$ & 1 & 1& 2& 3& 3\\ \hline $C_2$ & 1 & 1& 2& -1& -1\\ \hline $C_3$ & 1 & -1& 0& 1& -1\\ \hline $C_4$ & 1 & 1& -1& 0& 0\\ \hline $C_5$ & 1 & -1& 0&-1& 1\\ \hline \end{tabular} \caption{Characters of $S_4$} \label{tcars} \end{center} \end{table} For models of TB mixing, one starts from the $S_4$ presentation $A^3=B^4=(BA^2)^2=1$ and identifies, up to a similarity transformation, $B^2=S$ and $A=T$, where $S$ and $T$ are given in eqs. (\ref{trep}, \ref{ta4}). In this presentation one obtains a realisation of the 3-dimensional representation of $S_4$ where the $S$ and $A_{23}$ matrices in eq. (\ref{inv}) that leave invariant the TB form of $m_{\nu}$ in eq. (\ref{gl21}) as well as the matrix $T$ in eq. (\ref{ta4}) of invariance for $m_e^\dagger m_e$, all explicitly appear \cite{Bazzocchi:2009a}. In $S_4$ the $1'$ and $1''$ of $A_4$ are collected in a doublet. When the VEV of the doublet flavon is aligned along the $G_S$ preserving direction the resulting couplings are 2-3 symmetric as needed. In $A_4$ the 2-3 symmetry is only achieved if the $1'$ and $1''$ VEV's are identical (which is the $S_4$ prediction). As discussed in ref. \cite{Bazzocchi:2009a}, in the leptonic sector the main difference between $A_4$ and $S_4$ is that, while in the typical versions of $A_4$ the most general neutrino mass matrix depends on 2 complex parameters (related to the couplings of the singlet and triplet flavons), in $S_4$ it depends on 3 complex parameters (because the doublet is present in addition to singlet and triplet flavons). Other flavor groups have been considered for models of TB mixing. Some of them include $S_4$ as a subgroup, like $PSL_2(7)$ (the smallest group with complex triplet representations) \cite{Luhn:2007a, King:2009, King:2009a}, while others, like $\Delta(27)$ (which is a discrete subgroup of $SU(3)$) \cite{deMedeirosVarzielas:2007, Ma:2007b, Grimus:2007, Luhn:2007, Bazzocchi:2009c} or $Z_7\rtimes Z_3$ \cite{Luhn:2007b}, have no direct relation to $S_4$ \cite{King:2009b}. In Sect. 8 we will consider $S_4$ again in the different context of BM with large corrections from the lepton sector. A different approach to TB mixing has been proposed and developed in different versions by S. King and collaborators over the last few years \cite{King:2005, King:2006, deMedeirosVarzielas:2006, deMedeirosVarzielas:2007a, King:2009b}. The starting point is the decomposition of the neutrino mass matrix given in eqs. (\ref{1k1},\ref{4k1}) corresponding to exact TB mixing in the diagonal charged lepton basis: \begin{equation} \label{mLL} m_\nu= m_1\Phi_1 \Phi_1^T + m_2\Phi_2 \Phi_2^T + m_3\Phi_3 \Phi_3^T \end{equation} where $\Phi_1^T=\frac{1}{\sqrt{6}}(2,-1,-1)$, $\Phi_2^T=\frac{1}{\sqrt{3}}(1,1,1)$, $\Phi_3^T=\frac{1}{\sqrt{2}}(0,-1,1)$, are the respective columns of $U_{TB}$ and $m_i$ are the neutrino mass eigenvalues. Such decomposition is purely kinematical and does not possess any dynamical or symmetry content. In the King models the idea is that the three columns of $U_{TB}$ $\Phi_i$ are promoted to flavon fields whose VEVs break the family symmetry, with the particular vacuum alignments along the directions $\Phi_i$. Eq. (\ref{mLL}) directly arises in the see-saw mechanism, $m_\nu=m_D^T M^{-1} m_D$, written in the diagonal RH neutrino mass basis, $M={\rm diag}(M_1, M_2, M_3)$ when the Dirac mass matrix is given by $m_D^T=(v_1 \Phi_1,v_2 \Phi_2,v_3 \Phi_3)$, where $v_i$ are mass parameters describing the size of the VEVs. In this way, to each RH neutrino eigenvalue $M_i$, a particular light neutrino mass $m_i$ is associated. In the case of a strong neutrino hierarchy this idea can be combined with the framework of "Sequential Dominance", where the lightest RH neutrino, with its symmetry properties fixes the heaviest light neutrino and so on. For no pronounced hierarchy the correspondence between $M_i$ and $m_i$ can still hold and one talks of "Form Dominance" \cite{Chen:2009}. In these models the underlying family symmetry of the Lagrangian $G_f$ is completely broken by the the combined action of the $\Phi_i$ VEV's, and the flavor symmetry of the neutrino mass matrix emerges entirely as an accidental residual symmetry of the quadratic form of eq. (\ref{mLL}) \cite{King:2009b}. The symmetry $G_f$ plays a less direct role and the name "Indirect Models" is used by the authors. \section{Extension to quarks and GUT's} Much attention has been devoted to the question whether models with TB mixing in the neutrino sector can be suitably extended to also successfully describe the observed pattern of quark mixings and masses and whether this more complete framework can be made compatible with (supersymmetric) SU(5) or SO(10) Grand Unification. For models with approximate TB mixing in the leptonic sector we first consider the extension to quarks without Grand Unification and then the more ambitious task of building grand unified models. In GUT models based on $SU(5)\otimes G_f$ or $SO(10)\otimes G_f$ \footnote {The Pati-Salam group $SU(4)\otimes SU(2)\otimes SU(2)$ has also been considered, for example in \cite{King:2007,Toorop:2010yh}}, where $G_f$ is a flavor group, clearly all fields in a whole representation of $SU(5)$ or $SO(10)$ must have the same transformation properties under $G_f$. This poses a strong constraint on the way quarks and leptons have to transform under $G_f$. \subsection{Extension to quarks without GUT's} The simplest attempts of directly extending models based on $A_4$ to quarks have not been satisfactory. At first sight the most appealing possibility is to adopt for quarks the same classification scheme under $A_4$ that one has used for leptons (see, for example, \cite{Ma:2001, Ma:2002, Babu:2003, Altarelli:2006}). Thus one tentatively assumes that LH quark doublets $Q$ transform as a triplet $3$, while the RH quarks $(u^c,d^c)$, $(c^c,s^c)$ and $(t^c,b^c)$ transform as $1$, $1''$ and $1'$, respectively. This leads to $V_u=V_d$ and to the identity matrix for $V_{CKM}=V_u^\dagger V_d$ in the lowest approximation. This at first appears as very promising: a LO approximation where neutrino mixing is TB and $V_{CKM}=1$ is a very good starting point. But there are some problems. First, the corrections to $V_{CKM}=1$ turn out to be strongly constrained by the leptonic sector, because lepton mixing angles are very close to the TB values, and, in the simplest models, this constraint leads to a too small $V_{us}$ (i.e. the Cabibbo angle is rather large in comparison to the allowed shifts from the TB mixing angles) \cite{Altarelli:2006}. Also in these models, the quark classification which leads to $V_{CKM}=1$ is not compatible with $A_4$ commuting with SU(5). An additional consequence of the above assignment is that the top quark mass would arise from a non-renormalizable dimension five operator. In that case, to reproduce the top mass, we need to compensate the cutoff suppression by some extra dynamical mechanism. Alternatively, we have to introduce a separate symmetry breaking parameter for the quark sector, sufficiently close to the cutoff scale. Due to this, larger discrete groups have been considered for the description of quarks. A particularly appealing set of models is based on the discrete group $T'$, the double covering group of $A_4$ \cite{Frampton:1995, Aranda:2000, Aranda:2000a, Carr:2000, Aranda:2007, Frampton:2007, Frampton:2009, Ding:2009,Feruglio:2007, Chen:2007}. As we see in Table 2 the representations of $T'$ are those of $A_4$ plus three independent doublets 2, $2'$ and $2''$. The doublets are interesting for the classification of the first two generations of quarks \cite{Pomarol:1996, Barbieri:1996, Barbieri:1997, Barbieri:1997a}. For example, in ref. \cite{Feruglio:2007} a viable description was obtained, i.e. in the leptonic sector the predictions of the $A_4$ model are reproduced, while the $T'$ symmetry plays an essential role for reproducing the pattern of quark mixing. But, again, the classification adopted in this model is not compatible with Grand Unification. \subsection{Extension to quarks within GUT's} As a result, the group $A_4$ was considered by many authors to be too limited to also describe quarks and to lead to a grand unified description. It has been recently shown \cite{Altarelli:2008} that this negative attitude is not justified and that it is actually possible to construct a viable model based on $A_4$ which leads to a grand unified theory (GUT) of quarks and leptons with TB mixing for leptons and with quark (and charged lepton) masses and mixings compatible with experiment. At the same time this model offers an example of an extra dimensional SU(5) GUT in which a description of all fermion masses and mixings is accomplished. The formulation of SU(5) in extra dimensions has the usual advantages of avoiding large Higgs representations to break SU(5) and of solving the doublet-triplet splitting problem. The choice of the transformation properties of the two Higgses $H_5$ and $H_{\overline{5}}$ has a special role in this model. They are chosen to transform as two different $A_4$ singlets $1$ and $1'$. As a consequence, mass terms for the Higgs colour triplets are not directly allowed and their masses are introduced by orbifolding, \`{a} la Kawamura \cite{Witten:1985, Kawamura:2001, Faraggi:2001}. In this model, proton decay is dominated by gauge vector boson exchange giving rise to dimension six operators, while the usual contribution of dimension five operators is forbidden by the selection rules of the model. Given the large $M_{GUT}$ scale of SUSY models and the relatively huge theoretical uncertainties, the decay rate is within the present experimental limits. A see-saw realization in terms of an $A_4$ triplet of RH neutrinos $\nu^c$ ensures the correct ratio of light neutrino masses with respect to the GUT scale. In this model extra dimensional effects directly contribute to determine the flavor pattern, in that the two lightest tenplets $T_1$ and $T_2$ are in the bulk (with a doubling $T_i$ and $T'_i$, $i=1,2$ to ensure the correct zero mode spectrum), whereas the pentaplets $F$ and $T_3$ are on the brane. The hierarchy of quark and charged lepton masses and of quark mixings is determined by a combination of extra dimensional suppression factors and of $U(1)_{FN}$ charges, both of which only apply to the first two generations, while the neutrino mixing angles derive from $A_4$ in the usual way. If the extra dimensional suppression factors and the $U(1)_{FN}$ charges are switched off, only the third generation masses of quarks and charged leptons survive. Thus the charged fermion mass matrices are nearly empty in this limit (not much of $A_4$ effects remain) and the quark mixing angles are determined by the small corrections induced by those effects. The model is natural, since most of the small parameters in the observed pattern of masses and mixings as well as the necessary vacuum alignment are justified by the symmetries of the model. However, in this case, like in all models based on $U(1)_{FN}$, the number of $\mathcal{O}(1)$ parameters is larger than the number of measurable quantities, so that in the quark sector the model can only account for the orders of magnitude (measured in terms of powers of an expansion parameter) and not for the exact values of mass ratios and mixing angles. A moderate fine tuning is only needed to enhance the Cabibbo mixing angle between the first two generations, which would generically be of $\mathcal{O}(\lambda_C^2)$. The problem of constructing GUT models based on $SU(5)\otimes G_f$ or $SO(10)\otimes G_f$ with approximate TB mixing in the leptonic sector has been considered by many authors (see, for example \cite{Ma:2005b, Ma:2006a, Ma:2006b, Morisi:2007, Grimus:2008, Bazzocchi:2008b, Altarelli:2008, Ciafaloni:2009, Bazzocchi:2009,Antusch:2010es}, based on $A_4$). In our opinion most of the models are incomplete (for example, the crucial issue of VEV alignment is not really treated in depth as it should) and/or involve a number of unjustified, ad hoc fine tuning of parameters. An interesting model based on $SU(5)\otimes T'$ is discussed in ref. \cite{Chen:2007}. In this model the $SU(5)$ tenplets $T_3$ and $T_a$ ($a=1,2$) of the third and of the first two generations are classified as 1 and 2 of $T'$, respectively, while the $SU(5)$ pentaplets are in a 3 of $T'$. This model provides a good description of fermion masses and mixings and appears simpler than the model in ref. \cite{Altarelli:2008}, which is also based on $SU(5)$. However, the model of ref. \cite{Chen:2007} is fine tuned. In fact one does not understand how it is possible that, for example, the electron and the muon masses can come out so widely different as observed, given that in this model their left and right components separately transform in an identical way under $T'$. The reason is that in the second term of eq. 5 of ref. \cite{Chen:2007}, only one of three possible contractions has been taken into account. If the missing ones, which are also allowed by the assumed symmetry properties, are included with generic coefficients, one in fact finds that the $e$ and $\mu$ masses are of the same order in the absence of fine tuning. Given that the expansion parameter in the model is of $\mathcal{O}(\lambda_C)$ the fine tuning which is needed is large. One possible way out would be to invoke some ultraviolet completion of the model where particular heavy field exchanges could justify the presence of only the desired couplings after the heavy fields are integrated out. Also, in the model of ref. \cite{Chen:2007} there is no discussion of the origin of the required vacuum alignment. Recently some GUT models based on SU(5)$\times S_4$ have appeared \cite{Ishimori:2010xk,Hagedorn:2010th}. Also in these models the first two generation fermions are in the same $S_4$ representations (either a doublet, for tenplets, or a triplet, for pentaplets). In the absence of an additional principle the electron and muon mass should naturally be of the same order. In ref. \cite{Ishimori:2010xk} the vanishing of the electron mass at LO is obtained by the ad hoc choice of one particular minimum of the scalar potential among a continuous family of degenerate solutions (see their eqs. (70-71)). In the case of ref. \cite{Hagedorn:2010th} the problem is solved by introducing new heavy particles with suitable interactions that, once integrated out, produce the desired structure for the mass matrix. As for the models based on $SO(10)\otimes G_f$ we select two recent examples with $G_f=S_4$ \cite{Dutta:2009, Dutta:2009a} and $G_f=PSL_2(7)$ \cite{King:2009a}. Clearly the case of $SO(10)$ is even more difficult than that of $SU(5)$ because the neutrino sector is tightly related to that of quarks and charged leptons as all belong to the 16 of $SO(10)$ (for a general analysis of $SO(10)\otimes A_4$ see \cite{Bazzocchi:2008b}). The strategy adopted in refs. \cite{Dutta:2009, Dutta:2009a, King:2009a} as well as in other $SO(10)$ models, is as follows. One considers renormalisable fermion mass terms with Higgs multiplets of the $SO(10)$ 10 ($h$ terms) and 126 ($f$ terms) representations. The Majorana neutrino mass matrix arises from the 126. One assumes that the dominant contribution to the Dirac masses of fermions is from the $h$ terms with small corrections from the $f$ terms. In first approximation the $h$ contribution is a matrix of rank 1 with only the third generation mass being non vanishing. The light fermion masses and the quark mixings then arise from the $f$ terms (and from some possible extra terms). The third family dominance is obtained by a term with a double flavon factor in ref. \cite{Dutta:2009, Dutta:2009a} (based on $S_4$) which then makes particularly difficult to keep the corrective $f$ terms small (for this fine tuning is needed or a suitable ultraviolet completion). In ref. \cite{King:2009a} the dominant $h$ terms are induced by a single $PSL_2(7)$ sextet flavon (the existence of complex 3-dim and of 6-dim representations is the peculiarity of $PSL_2(7)$). In both models in the neutrino sector one has a sum of type I and II see-saw contributions of the form: \begin{equation} m_\nu=fv_L-m_D^T\frac{1}{fv_R}m_D\\ \label{I+II} \end{equation} where the first term is from the exchange of a triplet Higgs with VEV proportional to $v_L$ while the second term is from type I see-saw with RH mass proportional to $v_R$. One must assume that the first term is dominant and the second is negligible. Then the leading approximation for the fermion Dirac masses is from the $h$ terms and for neutrino masses from the $f$ terms. The $f$ terms are diagonalized by the TB mixing unitary matrix. In this way the connection between quarks and neutrinos is relaxed and a completely different pattern of mixing can be realized in the two sectors. Clearly for the $f v_L$ dominance in eq. (\ref{I+II}) one needs $v_L>>v^2/v_R$ with $v \sim h/f$ . This needs widely different scales for $v_L$ and $v_R$ in the model and much of the description of the corresponding dynamics, along the lines of refs. \cite{Goh:2004, Mohapatra:2007a}, remains to be studied in detail. In both of these models the discussion of the alignment is not satisfactory. In particular in ref. \cite{King:2009a} it is only proven that the arbitrary coefficients appearing in the most general allowed superpotential can be fitted to lead to the required ratios of components in the VEV's (while for a natural model one would require that the alignment automatically follows in a whole region of the parameter space). In conclusion, in our opinion, the problem of constructing a satisfactory natural model based on $SO(10)$ with built-in TB mixing at the LO approximation, remains open. \section{The $S_4$ group and BM mixing} If one takes the alternative view that the agreement with TB mixing is accidental and is rather oriented to consider "weak" complementarity as a more attractive guiding principle, then a better starting point could be BM mixing. In the BM scheme $\tan^2{\theta_{12}}= 1$, to be compared with the latest experimental determination: $\tan^2{\theta_{12}}= 0.45\pm 0.04$ (at $1\sigma$) \cite{Strumia:2006, Gonzalez-Garcia:2008a, Bandyopadhyay:2008, Fogli:2008, Fogli:2008a, Schwetz:2008, Maltoni:2008} (see fig.2), so that a rather large non leading correction is needed, as already mentioned. A discrete group approach can also work for BM mixing. We now summarise a model \cite{Altarelli:2009b} based on $S_4$ that leads to BM mixing in first approximation while the agreement with the data is restored by large NLO corrections that arise from the charged lepton sector. The group $S_4$ is particularly suitable for reproducing BM mixing in LO because the unitary matrices $S_{BM}$, given in eq. (\ref{matS}), and $T_{BM}$, given in eq. (\ref{ts4}), directly provide a presentation of $S_4$. We recall that $S_{BM}$ leaves invariant the most general mass matrix for BM mixing in the charged lepton diagonal basis, eq. (\ref{gl2}), while $T_{BM}$ leaves invariant the most general diagonal matrix $m_e^\dagger m_e$ for charged leptons (see eqs. (\ref{invS}, \ref{Tdiag})). In fact, from Table 2, we see that a possible presentation of $S_4$ is given by: \begin{equation} A^4=B^2=(AB)^3=1\\ \label{press4} \end{equation} In terms of 3x3 matrices, we can make the identifications $A=T_{BM}$ and $B=S_{BM}$ and eqs. (\ref{press4}) are satisfied. As was the case for the $A_4$ models, again in this model the invariance under $A_{23}$, which is also necessary to specify BM mixing according to eq. (\ref{invS}), arises accidentally as a consequence of the specific field content and is limited to the contribution of the dominant terms to the neutrino mass matrix. In the model the 3 generations of LH lepton doublets $l$ and of RH neutrinos $\nu^c$ to two triplets $3$, while the RH charged leptons $e^c$, $\mu^c$ and $\tau^c$ transform as $1$, $1'$ and $1$, respectively. The $S_4$ symmetry is then broken by suitable triplet flavons. All the flavon fields are singlets under the Standard Model gauge group. Additional symmetries are needed, as usual, to prevent unwanted couplings and to obtain a natural hierarchy among $m_e$, $m_\mu$ and $m_\tau$. The complete flavor symmetry of the model is $S_4\times Z_4\times U(1)_{FN}$. A flavon $\theta$, carrying a negative unit of the $U(1)_{FN}$ charge F, acquires a vacuum expectation value (VEV) and breaks $U(1)_{FN}$. A supersymmetric context is adopted, so that two Higgs doublets $h_{u,d}$, invariant under $S_4$, are present in the model as well as the $U(1)_R$ symmetry related to R-parity and the driving fields in the flavon superpotential. Supersymmetry also helps producing and maintaining the hierarchy $\langle h_{u,d}\rangle=v_{u,d}\ll \Lambda$ where $\Lambda$ is the cut-off scale of the theory. The fields in the model and their classification under the symmetry are summarized in Table \ref{table:TransformationsS}. The fields $\psi_l^0$, $\chi_l^0$ , $\xi_\nu^0$ and $\phi_\nu^0$ are the driving fields. \begin{table}[h] \begin{center} \begin{tabular}{|c||c|c|c|c|c|c||c||c|c|c|c||c|c|c|c|} \hline &&&&&&&&&&&&&&&\\[-0,3cm] & $l$ & $e^c$ & $\mu^c$ & $\tau^c$ & $\nu^c$ & $h_{u,d}$ & $\theta$ & $\phi_l$ & $\chi_l$ & $\psi_l^0$ & $\chi_l^0$ & $\xi_\nu$ &$\phi_\nu$ & $\xi_\nu^0$ & $\phi_\nu^0$ \\ &&&&&&&&&&&&&&&\\[-0,3cm] \hline &&&&&&&&&&&&&&&\\[-0,3cm] $S_4$ & 3 & 1 & $1^\prime$ & 1 & 3 & 1 & 1 & 3 & $3^\prime$ & 2 & $3'$ & 1 & 3 & 1 & 3 \\ &&&&&&&&&&&&&&&\\[-0,3cm] $Z_4$ & 1 & -1 & -i & -i & 1 & 1 & 1 & i & i & -1 & -1 & 1 & 1 & 1 & 1 \\ &&&&&&&&&&&&&&&\\[-0,3cm] $U(1)_{FN}$ & 0 & 2 & 1 & 0 & 0 & 0 & -1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\ &&&&&&&&&&&&&&&\\[-0,3cm] $U(1)_R$ & 1 & 1 & 1 & 1 & 1 & 1 & 0 & 0 & 0 & 2 & 2 & 0 & 0 & 2 & 2 \\ \hline \end{tabular} \end{center} \caption{\label{table:TransformationsS}Transformation properties of all the fields.} \end{table} The complete superpotential can be written as $w=w_l+w_\nu+w_d$. The $w_d$ term is responsible for the alignment. It was discussed in ref. \cite{Altarelli:2009b} and this discussion will not be repeated here. The terms $w_l$ and $w_\nu$ determine the lepton mass matrices (we indicate with $(\ldots)$ the singlet 1, with $(\ldots)^\prime$ the singlet $1^\prime$ and with $(\ldots)_V$ ($V=2,\,3,\,3'$) the representation V) \begin{eqnarray} w_l\;=&&\frac{y_e^{(1)}}{\Lambda^2}\frac{\theta^2}{\Lambda^2}e^c(l\phi_l\phi_l)+ \frac{y_e^{(2)}}{\Lambda^2}\frac{\theta^2}{\Lambda^2}e^c(l\chi_l\chi_l)+ \frac{y_e^{(3)}}{\Lambda^2}\frac{\theta^2}{\Lambda^2}e^c(l\phi_l\chi_l)+\nonumber\\ &+&\frac{y_\mu}{\Lambda}\frac{\theta}{\Lambda}\mu^c(l\chi_l)^\prime+\frac{y_\tau}{\Lambda}\tau^c(l\phi_l)+\dots \label{wl}\\ \nonumber\\ w_\nu\;=&&y(\nu^cl)+M \Lambda (\nu^c\nu^c)+a(\nu^c\nu^c\xi_\nu)+b(\nu^c\nu^c\phi_\nu)+\dots\\ \label{wd}\nonumber \end{eqnarray} where $a$ and $b$ are complex coefficients. Again, to keep our formulae compact, we omit to write the Higgs fields $h_{u,d}$. For instance $y_\tau \tau^c(l\phi_l)/\Lambda$ stands for $y_\tau \tau^c(l\phi_l)h_d/\Lambda$, $y(\nu^cl)$ stands for $y(\nu^cl) h_u$. The powers of the cutoff $\Lambda$ also take into account the presence of the omitted Higgs fields. Note that the parameters $M$, $M_\phi$, $M_\xi$ and $M'_\xi$ defined above are dimensionless. In the above expression for the superpotential $w$, only the lowest order operators in an expansion in powers of $1/\Lambda$ are explicitly shown. Dots stand for higher dimensional operators that will be discussed later on. The stated symmetries ensure that, for the leading terms, the flavons that appear in $w_l$ cannot contribute to $w_\nu$ and viceversa. The potential corresponding to $w_d$ possesses an isolated minimum for the following VEV configuration: \begin{equation} \displaystyle\frac{\mean{\phi_l}}{\Lambda}=\left( \begin{array}{c} 0 \\ 1 \\ 0 \\ \end{array} \right)A\qquad\qquad \displaystyle\frac{\mean{\chi_l}}{\Lambda}=\left( \begin{array}{c} 0 \\ 0 \\ 1 \\ \end{array} \right)B \label{vev:charged:best} \end{equation} \begin{equation} \hspace{-1.5cm} \displaystyle\frac{\mean{\phi_\nu}}{\Lambda}=\left( \begin{array}{c} 0 \\ 1 \\ -1 \\ \end{array} \right)C\qquad\quad \displaystyle\frac{\mean{\xi_\nu}}{\Lambda}=D \label{vev:neutrinos} \end{equation} where the factors $A$, $B$, $C$, $D$ should obey to the relations: \begin{eqnarray} &\sqrt{3}f_1A^2+\sqrt{3}f_2B^2+f_3AB=0 \label{AB}\\ \nonumber\\ &D=-\displaystyle\frac{M_\phi}{g_2}\qquad\qquad C^2=\displaystyle\frac{g_2^2M_\xi^2+g_3M_\phi^2-g_2M_\phi M'_\xi}{2 g_2^2g_4} \label{CD}\;. \end{eqnarray} Similarly, the Froggatt-Nielsen flavon $\theta$ gets a VEV, determined by the D-term associated to the local $U(1)_{FN}$ symmetry, and it is denoted by \begin{equation} \frac{\mean{\theta}}{\Lambda}= t\;. \label{deft} \end{equation} With this VEV's configuration, the charged lepton mass matrix is diagonal \begin{equation} m_l=\left( \begin{array}{ccc} (y_e^{(1)}B^2-y_e^{(2)}A^2+y_e^{(3)}AB)t^2 & 0 & 0 \\ 0 & y_\mu Bt & 0 \\ 0 & 0 & y_\tau A \\ \end{array} \right) v_d \end{equation} so that at LO there is no contribution to the $U_{PMNS}$ mixing matrix from the diagonalization of charged lepton masses. In the neutrino sector for the Dirac and RH Majorana matrices we have \begin{equation} m_\nu^D=\left( \begin{array}{ccc} 1 & 0 & 0 \\ 0 & 0 & 1 \\ 0 & 1 & 0 \\ \end{array} \right)yv_u\qquad\qquad M_N=\left( \begin{array}{ccc} 2M+2aD & -2bC & -2bC \\ -2bC & 0 & 2M+2aD \\ -2bC & 2M+2aD & 0 \\ \end{array} \right)\Lambda\;. \label{Feq:RHnu:masses} \end{equation} The matrix $M_N$ can be diagonalized by the BM mixing matrix $U_{BM}$, which represents the full lepton mixing at the LO, and the eigenvalues are \begin{equation} M_1=2|M+aD-\sqrt{2}bC|\Lambda\qquad M_2=2|M+aD+\sqrt{2}bC|\Lambda\qquad M_3=2|M+aD|\Lambda\;. \end{equation} After see-saw, since the Dirac neutrino mass matrix commutes with $M_N$ and its square is a matrix proportional to unity, the light neutrino Majorana mass matrix, given by the see-saw relation \mbox{$m_\nu=(m_\nu^D)^TM_N^{-1}m_\nu^D$}, is also diagonalized by the BM mixing matrix and the eigenvalues are \begin{equation} |m_1|=\frac{|y^2|v_u^2}{2|M+aD-\sqrt{2}bC|}\displaystyle\frac{1}{\Lambda}\qquad |m_2|=\frac{|y^2|v_u^2}{2|M+aD+\sqrt{2}bC|}\displaystyle\frac{1}{\Lambda}\qquad |m_3|=\frac{|y^2|v_u^2}{2|M+aD|}\displaystyle\frac{1}{\Lambda}\;. \label{spec} \end{equation} The light neutrino mass matrix depends on only 2 effective parameters, at LO, indeed the terms $M$ and $aD$ enter the mass matrix in the combination $F\equiv M+a D$. The coefficients $y_e^{(i)}$, $y_\mu$, $y_\tau$, $y$, $a$ and $b$ are all expected to be of $\mathcal{O}(1)$. A priori $M$ could be of $\mathcal{O}(1)$, corresponding to a RH neutrino Majorana mass of $\mathcal{O}(\Lambda)$, but, actually, it must be of the same order as $C$ and $D$. In the context of a grand unified theory this would correspond to the requirement that $M$ is of $\mathcal{O}(M_{GUT})$ rather than of $\mathcal{O}(M_{Planck})$. We expect a common order of magnitude for the VEV's (scaled by the cutoff $\Lambda$): \begin{equation} A \sim B \sim v\;,~~~~~~~~~~C \sim D \sim v'\;. \end{equation} However, due to the different minimization conditions that determine $(A,B)$ and $(C,D)$, we may tolerate a moderate hierarchy between $v$ and $v'$. Similarly the order of magnitude of $t$ is in principle unrelated to those of $v$ and $v'$. It is possible to estimate the values of $v$ and $t$ by looking at the mass ratios of charged leptons (while $v'$ only enters in the neutrino sector): and the result is that $t \sim 0.06$ and $v \sim 0.08$ (modulo coefficients of $\mathcal{O}(1)$). So far we have shown that, at LO, we have diagonal and hierarchical charged leptons together with the exact BM mixing for neutrinos. It is clear that substantial NLO corrections are needed to bring the model to agree with the data on $\theta_{12}$. A crucial feature of the model is that the neutrino sector flavons $\phi_\nu$ and $\xi_\nu$ are invariant under $Z_4$ which is not the case for the charged lepton sector flavons $\phi_l$ and $\chi_l$. The consequence is that $\phi_\nu$ and $\xi_\nu$ can contribute at NLO to the corrections in the charged lepton sector, while at NLO $\phi_l$ and $\chi_l$ cannot modify the neutrino sector couplings. As a results the dominant corrections to the BM mixing matrix only occur at NLO through the diagonalization of the charged leptons. In fact, at NLO the neutrino mass matrix is still diagonalized by $U_{BM}$ but the mass matrix of charged leptons is no more diagonal. Including these additional terms from the diagonalization of charged leptons the $U_{PMNS}$ matrix can be written as \begin{equation} U_{PMNS}=U_l^\dag U_{BM}\;, \end{equation} and therefore the corrections from $U_l$ affect the neutrino mixing angles at NLO according to \beq\begin{array}{l} \sin^2\theta_{12}=\displaystyle\frac{1}{2}-\frac{1}{\sqrt{2}}(V_{12}+V_{13})v'\\[0.2cm] \sin^2\theta_{23}=\displaystyle\frac{1}{2}\\[0.2cm] \sin\theta_{13}=\displaystyle\frac{1}{\sqrt{2}}(V_{12}-V_{13})v'\;. \label{sinNLO} \end{array}\eeq where the coefficients $V_{ij}$ arise from $U_l$. By comparing these expressions with the current experimental values of the mixing angles in Table 1, we see that, to correctly reproduce $\theta_{12}$ we need a parameter $v'$ of the order of the Cabibbo angle $\lambda_C$. Moreover, barring cancellations of/among some the $V_{ij}$ coefficients, also $\theta_{13}$ is corrected by a similar amount, while $\theta_{23}$ is unaffected at the NLO. A salient feature of this model is that, at NLO accuracy, the large corrections of $\mathcal{O}(\lambda_C)$ only apply to $\theta_{12}$ and $\theta_{13}$ while $\theta_{23}$ is unchanged at this order. As a correction of $\mathcal{O}(\lambda_C)$ to $\theta_{23}$ is hardly compatible with the present data (see Table 1) this feature is very crucial for the phenomenological success of this model. It is easy to see that this essential property depends on the selection in the neutrino sector of flavons $\xi_\nu$ and $\phi_\nu$ that transform as 1 and 3 of $S_4$, respectively. If, for example, the singlet $\xi_\nu$ is replaced by a doublet $\psi_\nu$ (and correspondingly the singlet driving field $\xi_\nu^0$ is replaced by a doublet $\psi_\nu^0$), all other quantum numbers being the same, one can construct a variant of the model along similar lines, but, in this case, all the 3 mixing angles are corrected by terms of the same order. This confirms that a particular set of $S_4$ breaking flavons is needed in order to preserve $\theta_{23}$ from taking as large corrections as the other two mixing angles. All this discussion applies at the NLO and we expect that at the NNLO the value of $\theta_{23}$ will eventually be modified with deviations of about $\mathcal{O}(\lambda^2_C)$. The next generation of experiments, in particular those exploiting a high intensity neutrino beam, will probably reduce the experimental error on $\theta_{23}$ and the sensitivity on $\theta_{13}$ to few degrees. All quantitative estimates are clearly affected by large uncertainties due to the presence of unknown parameters of order one, but in this model a value of $\theta_{13}$ much smaller than the present upper bound would be unnatural. If in the forthcoming generation of experiments no significant deviations from zero of $\theta_{13}$ will be detected, this construction will be strongly disfavoured. \section{Lepton flavor violation} Neutrino oscillations provide evidence of flavor conversion in the lepton sector. This indicates that lepton flavor violation (LFV) might take place, at least at some level, also in other processes such as those involving charged leptons. Flavor violating decays of charged leptons, strictly forbidden in the SM, are indeed allowed as soon as neutrino mass terms are considered. If neutrino masses are the only source of LFV, the effects are too small to be detected, but in most extensions of the SM where new particles and new interactions with a characteristic scale $\Lambda_{NP}$ are included, the presence of new sources of flavor violation, in both quark and lepton sectors, is a generic feature. The scale $\Lambda_{NP}$ can be much smaller than the cut-off scale $\Lambda$ introduced before. Indeed there are several indications suggesting new physics at the TeV scale, such as a successful gauge coupling unification, viable solutions to the hierarchy problem and realistic dark matter candidates. In a low-energy description, the associated effects can be parametrized by higher-dimensional operators. The dominant terms are represented by dimension six operators, suppressed by two powers of $\Lambda_{NP}$: \begin{equation} {\cal L}_{eff}= i\frac{e}{\Lambda_{NP}^2} {e^c}_i H^\dagger \sigma^{\mu\nu} F_{\mu\nu} {\cal Z}_{ij} l_j +\displaystyle\frac{1}{\Lambda_{NP}^2}[\tt 4-fermion~~ operators]+h.c. \label{leff} \end{equation} where $e$ is the electric charge and ${\cal Z}_{ij}$ denotes an adimensional complex matrix with indices in flavor space. If the underlying theory is weakly interacting with a typical coupling constant $g_{NP}$ and predicts new particles of mass $m_{NP}$ we expect $\Lambda_{NP}\approx 4\pi m_{NP}/g_{NP}$. The present bounds on the branching ratios \cite{Raidal:2008} of the rare charged lepton decays set stringent limits on combinations of the scale $\Lambda_{NP}$ and the coefficients of the involved operators. For instance, from $BR(\mu\to e \gamma)<1.2\times 10^{-11}$ \cite{Brooks:1999,Adam:2009} we get $|{\cal Z}_{\mu e}|<10^{-8}\times [\Lambda_{NP}({\rm TeV})/1~{\rm TeV}]^2$. Typically, for coefficients of order one, the existing bounds require a large scale $\Lambda_{NP}$, several orders of magnitude larger than the TeV scale. Conversely, to allow for new physics close to the TeV scale, coefficients much smaller than one are required, which may indicate the effect of a flavor symmetry. In theories with a flavor symmetry group $G_f$ spontaneously broken by a set of small parameters $\varepsilon$, the coefficients of the effective lagrangian in eq. (\ref{leff}) become functions of $\varepsilon$. The low-energy Lagrangian of eq. (\ref{leff}) is derived from the theory defined close to the cut-off scale $\Lambda$, where all operators are invariant under $G_f$ thanks to their dependence on the flavon multiplets. Below the flavor symmetry breaking scale the flavons are replaced by their VEVs, which enter the coefficients of ${\cal L}_{eff}$ through the dimensionless combination $\varepsilon\approx VEV/\Lambda$. Exploiting the smallness of the parameters $\varepsilon$ we can keep in ${\cal L}_{eff}$ the first few terms of a power series expansion. For instance: \begin{equation} {\cal Z}_{ij}\equiv{\cal Z}_{ij}\left(\varepsilon\right)={\cal Z}_{ij}^{(0)}+{\cal Z}_{ij}^{(1)}~\varepsilon+{\cal Z}_{ij}^{(2)}~\varepsilon^2... \end{equation} Notice that the same symmetry breaking parameters that control lepton masses and mixing angles also control the flavor pattern of the operators in ${\cal L}_{eff}$. This result is interesting in several respects. First of all the presence of the factors $\varepsilon^n$ can help in suppressing the rates of rare charged lepton decays while allowing for a relatively small and accessible scale $\Lambda_{NP}$. Second, once the above expansion has been determined in a given model, it could be possible to establish characteristic relations among LFV processes as a consequence of flavor symmetries and of their pattern of symmetry breaking. Finally, if $\Lambda_{NP}$ is sufficiently small, this opens the possibility that new particles might be produced and detected at the LHC, with features that could additionally confirm or reject the assumed symmetry pattern. All this allows, at least in principle, to realize an independent test of the flavor symmetry in the charged lepton sector. While the size of the scale $\Lambda_{NP}$ could be relatively small, in our presentation we assume that the flavour scale or cutoff $\Lambda$ is extremely large, possibly as large as the GUT scale. Then all low-energy effects due to the flavon dynamics are essentially those associated to their VEVs, which enter the effective higher dimensional operators through the dimensionless combination $\epsilon$. Virtual flavon exchanges give rise to other higher dimensional operators which are depleted by inverse power of $\Lambda$ and can be safely neglected. A much richer variety of effects due to the flavour dynamics would be possible if the scale $\Lambda$ were much smaller, close to the 100 TeV energy range, but we do not consider this possibility here. The effects described by ${\cal L}_{eff}$ are well-known. In a field basis where the kinetic terms are canonical and the charged lepton mass matrix is diagonal the real and imaginary parts of the diagonal matrix elements ${\cal Z}_{ii}$ are proportional to the anomalous magnetic moments (MDM) $a_i$ and to the electric dipole moments (EDM) $d_i$ of charged leptons, respectively: \begin{equation} a_i=2 m_i^{ch} \frac{v}{\sqrt{2} \Lambda_{NP}^2}Re {\cal Z}_{ii}~~~,~~~~~~~d_i=e \frac{v}{\sqrt{2} \Lambda_{NP}^2}Im {\cal Z}_{ii}~~~. \label{dm} \end{equation} The off-diagonal elements ${\cal Z}_{ij}$ describe the amplitudes for the radiative decays of the charged leptons: \begin{equation} R_{ij}=\frac{BR(l_i\to l_j\gamma)}{BR(l_i\to l_j\nu_i{\bar \nu_j})}=\frac{12\sqrt{2}\pi^3 \alpha}{G_F^3 {m_i^{\tt ch}}^2 \Lambda_{NP}^4}\left(\vert{\cal Z}_{ij}\vert^2+\vert{\cal Z}_{ji}\vert^2\right) \label{dt} \end{equation} where $\alpha$ is the fine structure constant, $G_F$ is the Fermi constant and $m_i^{ch}$ is the mass of the lepton $l_i$. Finally the four-fermion operators, together with the dipole operators controlled by ${\cal Z}$, describe other flavor violating processes like $\mu\to eee$, $\tau\to\mu\mu\mu$, $\tau\to eee$. An interesting example of flavor symmetry is that of minimal flavor violation (MFV) \cite{Chivukula:1987,Hall:1990,Ciuchini:1998,Buras:2001,D'Ambrosio:2002,Cirigliano:2005,Cirigliano:2006,Davidson:2006,Cirigliano:2007} whose (minimal) flavor symmetry group in the lepton sector is $G_f=SU(3)_{e^c}\times SU(3)_l$. Electroweak singlets $e^c$ and doublets $l$ transform as $(3,1)$ and $(1,\bar{3})$, respectively. The flavon fields or, better, their VEVs are the Yukawa couplings of the charged leptons, $Y_l=m_l/v$, and the adimensional coupling constants $\eta$ of the five-dimensional operator $O_5$ in eq. (\ref{O5}). They transform as $(\bar{3},3)$ and $(1,6)$, respectively. In a basis where the charged leptons are diagonal, we have \begin{equation} Y_l=\frac{\sqrt{2}}{v} m_l^{\rm diag}~~~,~~~~~~~~~~\eta=\frac{M}{v^2} U^* m_\nu^{\rm diag} U^\dagger~~~, \end{equation} where $M$ here denotes the mass scale suppressing the operator $O_5$. In MFV models the leading off-diagonal elements of ${\cal Z}_{ij}$ are given by: \begin{eqnarray} {\cal Z}_{ij}&=&c~ (Y_l~ \eta^\dagger \eta)_{ij}\nonumber\\ &=&\sqrt{2}c~\frac{m_i^{\tt ch}}{v}\frac{M^2}{v^4}\left[\Delta m^2_{sol} U_{i2} U^*_{j2}\pm\Delta m^2_{atm} U_{i3} U^*_{j3}\right] \label{zij} \end{eqnarray} where $c$ is an overall coefficient of order one and the plus (minus) sign refers to the case of normal (inverted) hierarchy. We see that, due to the presence of the ratio $M^2/v^2$ the overall scale of these matrix elements is poorly constrained. This is due to the fact that MFV does not restrict the overall strength of the coupling constants $\eta$, apart from the requirement that they remain in the perturbative regime. Very small or relatively large (but smaller than one) $\eta$ can be accommodated by adjusting the scale $M$. Thus, even after fixing $\Lambda_{NP}$ close to the TeV scale, in MFV the non-observation of $l_i\to l_j \gamma$ could be justified by choosing a small $M$, while a positive signal in $\mu\to e \gamma$ with a branching ratio in the range $1.2\times 10^{-11}\div 10^{-13}$ could also be fitted by an appropriate $M$, apart from a small region of the $\theta_{13}$ angle, around $\theta_{13}\approx0.02$ where a cancellation can take place in the left-hand side of eq. (\ref{zij}). The dependence on the scales $M$ and $\Lambda_{NP}$ can be eliminated by considering ratios of branching ratios. For instance: \begin{equation} \displaystyle\frac{R_{\mu e}}{R_{\tau \mu}}= \left\vert\frac{2\Delta m^2_{sol}}{3\Delta m^2_{atm}}\pm \sqrt{2}\sin\theta_{13} e^{i\delta}\right\vert^2<1~~~, \label{mfv} \end{equation} where we took the TB ansatz to fix $\theta_{12}$ and $\theta_{23}$. We see that $BR(\mu\to e\gamma)<BR(\tau\to \mu\gamma)$ always in MFV. Moreover, for $\theta_{13}$ above approximately $0.07$, $BR(\mu\to e\gamma)<1.2\times 10^{-11}$ implies $BR(\tau\to \mu\gamma)<10^{-9}$. For $\theta_{13}$ below $0.07$, apart possibly from a small region around $\theta_{13}\approx0.02$, both the transitions $\mu\to e \gamma$ and $\tau\to\mu\gamma$ might be above the sensitivity of the future experiments. The present limits are $BR(\tau\to\mu\gamma)<1.6\times 10^{-8}$ and $BR(\tau\to e\gamma)<9.4\times 10^{-8}$. A future super B factory might improve them by about one order of magnitude. In the SUSY case there are two doublets in the low-energy Lagrangian and we should take into account the $\tan\beta$ dependence. A different result for the matrix ${\cal Z}$ is obtained in the model described in Sect. 4 where $G_f=A_4\times Z_3\times U(1)_{FN}$. Starting from the relevant set of invariant operators, after the breaking of the flavor and electroweak symmetries, and after moving to a basis with canonical kinetic terms and diagonal mass matrix for charged leptons, we find \cite{Feruglio:2009,Feruglio:2008}: \begin{equation} \mathcal{Z} = \left( \begin{array}{ccc} \mathcal{O}(t^2 \varepsilon) & \mathcal{O}(t^2 \varepsilon^2) & \mathcal{O}(t^2 \varepsilon^2)\\ \mathcal{O}(t \varepsilon^2) & \mathcal{O}(t \varepsilon) & \mathcal{O}(t \varepsilon^2)\\ \mathcal{O}(\varepsilon^2) & \mathcal{O}(\varepsilon^2) & \mathcal{O}(\varepsilon) \end{array} \right) \label{hatM} \end{equation} where each matrix element is known only up to an unknown order-one dimensionless coefficient. There are two independent symmetry breaking parameters. The parameter $t=\langle\theta\rangle/\Lambda$ controls the charged lepton mass hierarchy and $\varepsilon= v_T/\Lambda$ describes the breaking of $A_4$. Notice that the uncertainty in the overall scale of the matrix elements ${\cal Z}_{ij}$ is related to the parameter $\varepsilon$ and is much smaller than the corresponding uncertainty in MFV. We can see that MDMs and EDMs arise at the first order in the parameter $\varepsilon$. By assuming that the unknown coefficients have absolute values and phases of order one, from eqs. (\ref{dm}) and (\ref{hatM}) we have: \begin{equation} a_i=\mathcal{O}\left(2\displaystyle\frac{{m_i^{\tt ch}}^2}{\Lambda_{NP}^2}\right)~~~,~~~~~~~~~~d_i=\mathcal{O}\left(e\displaystyle\frac{m_i^{\tt ch}}{\Lambda_{NP}^2}\right)~~~. \label{oom} \end{equation} From the existing limits on MDMs and EDMs and by using eqs. (\ref{oom}) as exact equalities we find the results shown in table 7. \begin{table}[!ht] \centering \begin{tabular}{|c|c|} \hline & \\[-9pt] $d_e<1.6\times 10^{-27}~~e~cm$&$\Lambda_{NP}>80~~{\rm TeV}$\\[3pt] \hline &\\[-9pt] $d_\mu<2.8\times 10^{-19}~~e~cm$&$\Lambda_{NP}>80~~{\rm GeV}$\\[3pt] \hline &\\[-9pt] $\delta a_e<3.8\times 10^{-12}$&$\Lambda_{NP}>350~~{\rm GeV}$\\[3pt] \hline &\\[-9pt] $\delta a_\mu= 302\pm88\times 10^{-11}$&$\Lambda_{NP}\approx 2.7~~{\rm TeV}$\\[3pt] \hline \end{tabular} \caption{Experimental limits on lepton MDMs and EDMs and corresponding bounds on the scale $\Lambda_{NP}$, derived from eq. (\ref{oom}). The data on the $\tau$ lepton have not been reported since they are much less constraining. For the anomalous magnetic moment of the muon, $\delta a_\mu$ stands for the deviation of the experimental central value from the SM expectation \cite{Bennett:2004,Passera:2009}.} \end{table} \vskip 0.2cm \noindent Concerning the flavor violating dipole transitions, from eq. ({\ref{hatM}) we see that the dominant contribution to the rate for $l_i\to l_j\gamma$ is given by: \begin{equation} \frac{BR(l_i\to l_j\gamma)}{BR(l_i\to l_j\nu_i{\bar \nu_j})}=\frac{48\pi^3 \alpha}{G_F^2 \Lambda_{NP}^4}\vert w_{ij} ~\varepsilon\vert^2 \label{LFV} \end{equation} where $w_{ij}$ are numbers of order one. As a consequence, the branching ratios of the three transitions $\mu\to e\gamma $, $\tau\to\mu\gamma$ and $\tau\to e\gamma$ are all expected be of the same order: \begin{equation} BR(\mu\to e \gamma)\approx BR(\tau\to\mu\gamma)\approx BR(\tau\to e \gamma)~~~. \label{equalbr} \end{equation} This is a distinctive feature of this class of models. Given the present experimental bound on $BR(\mu\to e \gamma)$, eq. (\ref{equalbr}) implies that $\tau\to\mu\gamma$ and $\tau\to e \gamma$ have rates much below the present and expected future sensitivity. Moreover, from the current (future) experimental limit on $BR(\mu\to e \gamma)$ \cite{Brooks:1999,Adam:2009,Meg} and assuming $\vert w_{\mu e}\vert=1$, we derive the following bound on $\vert \varepsilon/\Lambda_{NP}^2\vert$: \begin{equation} BR(\mu\to e \gamma)<1.2\times 10^{-11}~(10^{-13})~~~~~~~ \left\vert\displaystyle\frac{\varepsilon}{\Lambda_{NP}^2}\right\vert<1.2\times 10^{-11}~(1.1\times 10^{-12})~~{\rm GeV}^{-2}~~~. \end{equation} \noindent Taking two extreme values for the parameter $\vert \varepsilon\vert$ we find \begin{eqnarray} \Lambda_{NP}>20~(67)~~{\rm TeV}~&[\vert \varepsilon\vert = 0.005]\nonumber\\ \Lambda_{NP}>65~(210)~~{\rm TeV}&~~~[\vert \varepsilon\vert = 0.05]~~~. \end{eqnarray} This model also allows for four-fermion operators that are not suppressed by any power of the small parameter $t$ or $\varepsilon$ and that violate the individual lepton numbers $L_i$ \cite{Feruglio:2010qu}. They are all characterized by the selection rule $\Delta L_e \Delta L_\mu \Delta L_\tau=2$. For instance, one such operator is \begin{equation} ({\bar l} l)'({\bar l} l)''= \left[{\bar l}_e l_\tau {\bar l}_\mu l_\tau+ {\bar l}_\mu l_e {\bar l}_\tau l_e+ {\bar l}_\tau l_\mu {\bar l}_e l_\mu+h.c.\right]+...~~~. \end{equation} where dots stand for additional flavor conserving contributions. These operators can contribute to LFV decays such as $\tau^-\to \mu^+ e^- e^-$, $\tau^-\to e^+ \mu^- \mu^-$ and their conjugate, whose branching ratios have upper bounds of the order of $10^{-7}$\cite{Amsler:2008}. Through a rough dimensional estimate we find a lower bound on the scale $\Lambda_{NP}$ of the order of $15$ TeV. From the previous considerations we see that, even invoking a cancellation in the imaginary part of ${\cal{Z}}_{ee}$ to suppress the contribution to the electron EDM, it is difficult to avoid the conclusion that the scale $\Lambda_{NP}$ should lie considerably above the TeV range. We recall that if the operator in eq. (\ref{leff}) originates from one-loop diagrams via the exchange of weakly interacting particles of masses $m_{NP}$, then in our normalization a lower bound on $\Lambda_{NP}$ of 20 TeV corresponds to a lower bound on $m_{NP}$ of about $g_{NP}\Lambda_{NP}/(4\pi)\approx1~$ TeV, assuming $g_{NP}$ similar to the SU(2) gauge coupling. All the previous estimates are based on an effective Lagrangian approach, with no explicit reference to the dynamics at the scale $\Lambda_{NP}$. If the degrees of freedom associated to the new physics at the scale $\Lambda_{NP}$ and their interactions are known, it is possible to directly compute the amplitudes of interest. For instance, the SUSY model of Sect. 4 can be completed by adding a set of soft SUSY breaking terms, which are constrained by the invariance under $G_f=A_4\times Z_3\times U(1)_{FN}$ and its pattern of symmetry breaking \cite{Ishimori:2008,Hayakawa:2009,feruglio:2009a,Ding:2009b}. LFV amplitudes arise at one-loop level, via exchange of sleptons, charginos and neutralinos with masses of order $m_{SUSY}$. An explicit computation of $BR(l_i\to l_j\gamma)$ confirms both the predictions of eq. (\ref{equalbr}) and the behaviour of eq. (\ref{LFV}), with $\Lambda_{NP}=(4\pi/g) m_{SUSY}$ . The coefficients $w_{ij}$ are typically of ${\cal O}(0.1)$. When $\varepsilon$ is small, which also entails small $\tan\beta$ in our model, relatively light SUSY particles are allowed, while for $\varepsilon$ close to its upper limit, 0.05, SUSY particle masses of several hundred GeV or close to the TeV are needed to satisfy the present bound on $BR(\mu\to e \gamma)$, particularly if $\tan\beta$ is larger than 10. In either case there is only a very limited region of the parameter space where it is possible to explain the observed discrepancy in the muon MDM and to satisfy at the same time the current limit on $BR(\mu\to e \gamma)$. An interesting special case is that of universal SUSY breaking terms, giving rise to a cancellation in the elements of ${\cal Z}_{ij}$ below the diagonal \cite{feruglio:2009b,feruglio:2009a}. Under these circumstances $BR(l_i\to l_j\gamma)$ scale as $\varepsilon^4$ rather than as $\varepsilon^2$, with the possibility of much lighter SUSY particles. In SUSY $A_4$ models also LFV 4-fermion operators are depleted by powers of $\varepsilon$ and the corresponding bounds on $m_{SUSY}$ are relaxed. In the model discussed in Sect. 8, with $G_f=S_4\times Z_4\times U(1)_{FN}$, the matrix ${\cal Z}$ is given by \cite{masiero:2009}: \begin{equation} \mathcal{Z} = \left( \begin{array}{ccc} \mathcal{O}(t^2 v^2) & \mathcal{O}(t^2 v^2 v') & \mathcal{O}(t^2 v^2 v')\\ \mathcal{O}(t v v') & \mathcal{O}(t v) & \mathcal{O}(t v v'^2)\\ \mathcal{O}(v v') & \mathcal{O}(v v'^2) & \mathcal{O}(v) \end{array} \right) \label{hatM2} \end{equation} Predictions for EDMs and MDMs and corresponding bounds are similar to those discussed above in the case of the $A_4$ model and summarized in Table 7. Concerning the radiative decays of the charged leptons we find that $R_{\mu e}$ and $R_{\tau e}$ scale as $v'^2/\Lambda_{NP}^4$, whereas $R_{\tau\mu}$ scales as $v'^4/\Lambda_{NP}^4$. In this case the symmetry breaking parameter $v'$ is considerably larger than the parameter $\varepsilon$ of the $A_4$ model and this gives rise to more restrictive bounds on the scale of new physics $\Lambda_{NP}$. From $BR(\mu\to e \gamma)<1.2\times 10^{-11}~(10^{-13})$ we get: \begin{eqnarray} \Lambda_{NP}>90~(300)~~{\rm TeV}~&[v' = 0.1]\nonumber\\ \Lambda_{NP}>130~(430)~~{\rm TeV}&~~~[v' = 0.2]~~~. \end{eqnarray} The model also predicts: \begin{equation} BR(\mu\to e \gamma)\approx BR(\tau\to e\gamma)\gg BR(\tau\to \mu \gamma)~~~. \end{equation} Summarizing, in models with discrete flavor symmetries LFV processes are generically suppressed by the presence of small symmetry breaking parameters. However such a suppression is not completely efficient, at least in the explored models, to guarantee a scale of new physics close to the TeV. The best case is the one of the $A_4$ model, thanks to the very small expansion parameter $\varepsilon$. In specific SUSY realizations of the $A_4$ symmetry the present limits on the branching ratios of LFV processes still allow for a relatively light spectrum of superparticles, in a region of masses of interest to LHC. \section{Leptogenesis} The violation of $B-L$ implied by the see-saw mechanism suggests an interesting link between neutrino physics and the mechanism that produced the observed baryon asymmetry in the early universe. If the baryon asymmetry originates well above the electroweak scale, $B-L$ violation represents a necessary condition, since any initial $B+L$ asymmetry would be erased in the subsequent evolution of the universe. According to leptogenesis the asymmetry is determined by the CP violating, out-of-equilibrium decays of the heavy RH neutrinos \cite{Fukugita:1986}. Through $B-L$ non-conservation of neutrino interactions, the asymmetry is first generated in the leptonic number and then partly converted into the observed baryonic one via sphaleron interactions. Depending on whether the relevant decays occur at a sufficiently high temperature or not, we have an unflavored regime, where the leptons in the final state are indistinguishable, or a flavored regime, where the specific interactions of the different leptons in the decay products cannot be neglected \cite{Abada:2006,Nardi:2006}. It is also quite remarkable that, at least in its simplest implementation, leptogenesis requires light neutrino masses below the eV scale \cite{Buchmuller:2003,Buchmuller:2004,Giudice:2004,Buchmuller:2005}, in a range which is fully compatible with other experimental constraints. Unfortunately, without any additional assumptions, it is difficult to promote this elegant picture into a testable theory, due to the large number of independent parameters of the see-saw model. Models of lepton masses based on flavor symmetries typically depend on a restricted number of parameters, thus opening the interesting possibility of relating the baryon asymmetry to other low-energy observables. As a general rule, to provide a realistic description of lepton masses and mixing angles the flavor symmetry should always be broken. The breaking is described by a set of small dimensionless quantities $\varepsilon$, which provide efficient expansion parameters. As we have seen in the previous sections, small observable quantities such as charged lepton mass ratios, $\theta_{13}$, $\theta_{23}-\pi/4$ can be expanded in power series of $\varepsilon$, and the predictions are dominated by the lowest (positive) power. In the context of leptogenesis, given the extreme smallness of the baryon asymmetry \cite{Komatsu:2009} \begin{equation}\label{etaBobs} \eta_B^{\rm CMB} = (6.2 \pm 0.15)\times 10^{-10} \, , \end{equation} it can be convenient, at least in a certain regime, that the $C\!P$ asymmetries in the RH neutrino decays are also suppressed by powers of $\varepsilon$. If the baryon asymmetry is dominated by the decay of a single RH neutrino, we can write \footnote{We will denote the $C\!P$ asymmetries with $\xi$ and we keep the letter $\varepsilon$ to indicate the generic expansion parameter of a spontaneously broken flavor symmetry.}: \begin{equation} \eta_B= d~ \xi~ k \end{equation} where $d$ describes the combined effect of sphaleron conversion and dilution from photon production, $\xi$ is the relevant $C\!P$ asymmetry and $k$ takes into account the wash-out effects. Typically we expect a dilution factor $d$ of order $10^{-2}$ and, barring fine-tuning of the parameters, a wash-out factor $k$ in the range $10^{-3}\div 10^{-2}$, which favors $\xi$ around $10^{-6}\div 10^{-5}$. Such $C\!P$ asymmetry arises from the interference of the tree-level and the one-loop decay amplitudes and depends quadratically on the neutrino Yukawa couplings. In models like the ones discussed in Sects. 4 and 8, where the RH neutrino masses are very large, close to $10^{14}$ GeV, and the corresponding neutrino Yukawa couplings are of $\mathcal{O}(1)$, a rough estimate of the total $C\!P$ asymmetry would give $\xi=\mathcal{O}(1/(8\pi))$, by far too large compared to $10^{-6}\div 10^{-5}$. It is therefore interesting to analyze under which conditions the $C\!P$ asymmetries vanish in the limit of exact symmetry, so that the first non-vanishing contribution is given by some power of the symmetry breaking parameters $\varepsilon$. If the $C\!P$ asymmetry relevant for leptogenesis is suppressed by powers of $\varepsilon$, this opens the very interesting possibility of relating the observed baryon asymmetry $\eta_B$ to other low-energy observable quantities \cite{Lin:2009a, Mohapatra:2005a, Mohapatra:2005b} such as $\theta_{13}$, $\theta_{23}-\pi/4$, $BR(l_i\to l_j \gamma)$. The total $C\!P$ asymmetries in the decay of a RH neutrino $\nu^c_i$ are \begin{equation} \xi_i=\displaystyle\frac{\Gamma_i-\overline{\Gamma}_i}{\Gamma_i+\overline{\Gamma}_i} \end{equation} where $\Gamma_i$ $(\overline{\Gamma}_i)$ is the decay rate of $\nu^c_i$ into leptons (antileptons). In the flavored regime the relevant asymmetries $\xi_{if}$ involve final states with a specific lepton flavor $f$. The flavored regime takes place for $M_i\le c~10^{12}$ GeV where $c=1$ $(1+\tan^2\beta)$ in the ordinary (SUSY) case. The unflavored regime occurs for RH neutrino masses above that threshold. At one-loop we have~\cite{Covi:1996}: \begin{eqnarray} \label{eq:flCPasymm} \xi_{if} &=& \frac{1}{8\pi \hat{\cal Y}_{ii}}\sum_{j\neq i} \left\{ {\rm Im}\left[ \hat{\cal Y}_{ij} \hat Y_{if} \hat Y^*_{jf}\right] f_{ij} + {\rm Im}\left[ \hat{\cal Y}_{ji} \hat Y_{if}\hat Y^*_{jf}\right] g_{ij}\right\} \\ \label{eq:CPasymm} \xi_{i} &=& \sum_f \xi_{if} = \frac{1}{8\pi \hat{\cal Y}_{ii}}\sum_{j\neq i} {\rm Im} \left[\hat{\cal Y}_{ij}^2\right] f_{ij} \, , \end{eqnarray} where ${\cal Y}$ is a combination of the neutrino Yukawa couplings $Y=m_\nu^D/v_u$ \begin{equation} \label{eq:YY} {\cal Y}_{ij} = \left(Y Y^\dagger\right)_{ij}~~~. \end{equation} and the hat in eqs. (\ref{eq:flCPasymm},\ref{eq:CPasymm}) denotes a basis where the mass matrix $M$ of heavy Majorana neutrinos and that of charged leptons, $m_l$, are diagonal. The functions $f_{ij}$ and $g_{ij}$ depend on the mass ratios of the RH neutrino masses $M_i$. From eqs. (\ref{eq:flCPasymm},\ref{eq:CPasymm}) we see that both $\xi_{if}$ and $\xi_i$ vanish if $\hat{\cal Y}$ is diagonal. The total asymmetries $\xi_i$ vanish also if $\hat{\cal Y}$ has real non-diagonal entries. A necessary and sufficient condition for a diagonal $\hat{\cal Y}$ is: \begin{equation} {{\cal Y}} M-M {{\cal Y}}^T =0~~~, \label{com1} \end{equation} where the matrices ${\cal Y}$ and $M$ are evaluated in any basis. If the model is invariant under the action of a flavor symmetry group $G_f$ we have an interesting sufficient condition for the vanishing of the $C\!P$ asymmetries. If the heavy RH neutrinos transform in a (three-dimensional) irreducible representation of $G_f$, then in the limit of exact symmetry, where the symmetry breaking parameters $\varepsilon$ go to zero, all $C\!P$ asymmetries vanish \cite{Bertuzzo:2009}. In this limit it is possible to show that ${\cal Y}$ becomes proportional to the unit matrix as a consequence of a completely general group theoretical property. Thus, from eqs. (\ref{eq:flCPasymm},\ref{eq:CPasymm}) we conclude that the asymmetries $\xi_i$ and $\xi_{if}$ vanish. Notice that irreducible representations of dimension larger than one are only possible if $G_f$ is non-abelian. Beyond the symmetry limit, in general $\hat{\cal Y}$ gets corrections and develops complex off-diagonal entries at some order $\varepsilon^p$. If the spectrum of RH neutrinos is non-degenerate in the symmetry limit, we expect $\xi_i={\cal O}(\varepsilon^{2p})$ and $\xi_{if}={\cal O}(\varepsilon^{p})$. Degeneracy of RH neutrinos can modify this behavior through the dependence on $\varepsilon$ of the functions $f_{ij}$ and $g_{ij}$. This result applies to both the models described in Sects. 4 and 8, where the RH neutrinos transform in the three-dimensional representions of $A_4$ and $S_4$, respectively. In the limit of exact flavor symmetry we find in both cases ${\cal Y}=|y|^2~{\bf 1}$ where ${\bf 1}$ denotes the identity matrix. This equality holds in any basis, in particular in the mass eigenstate basis of RH neutrinos and we have $\xi_i=0$ in the symmetry limit. In both models all RH neutrino are very heavy, with masses well above $10^{12}$ GeV, and the unflavored regime applies. In the $A_4$ model of Sect. 4, $\hat{\cal Y}$ acquires complex off-diagonal entries of order $\varepsilon\approx v_T/\Lambda$. The $C\!P$ asymmetries $\xi_i$ depend only on three real parameters: two independent real symmetry breaking parameters $\varepsilon_i$ and the lightest neutrino mass. In particular there is only one independent phase which is determined by the lightest neutrino mass up to an overall sign. We have approximately \cite{Jenkins:2008} \begin{equation} \xi_i\approx \displaystyle\frac{\varepsilon^2}{8\pi} \label{cpa} \end{equation} More precisely \cite{Bertuzzo:2009,Hagedorn:2009,Riva:2010}, for normal ordering of the neutrino mass spectrum all asymmetries $\xi_i$ are of the same order of magnitude. For inverted ordering the two asymmetries $\xi_{1,2}$ get enhanced compared to the approximate estimate of eq. (\ref{cpa}) by a factor $~10^{3}$ coming from the functions $f_{12}$ and $f_{21}$, as a result of the near degeneracy of two heavy RH neutrinos. To reproduce the observed baryon asymmetry, eq. (\ref{etaBobs}), different wash-out effects are required in the two cases. In the case of normal ordering the experimental value in eq. (\ref{etaBobs}) is obtained when the parameter $\varepsilon$ is in its natural window, $5\times 10^{-3}\div 5\times 10^{-2}$, for a wide range of neutrino Yukawa couplings $y$. For inverted ordering a much larger wash-out suppression is needed. When $\varepsilon$ falls in the optimal range $5\times 10^{-3}\div 5\times 10^{-2}$ this can be accommodated by restricting both $y\times\sin\beta$ and $m_3$ in a limited range. It is quite remarkable that in both cases the range of the symmetry breaking parameter $\varepsilon$ suggested by the constraints on lepton masses and mixing angles corresponds to that required to get the observed baryon asymmetry through leptogenesis. In the $S_4$ model discussed in Sect. 8, $\hat{\cal Y}$ acquires complex off-diagonal entries at the order $v^4/v'$ and the $C\!P$ asymmetries are expected to be of order $v^8/(v'^2 8\pi)$. Assuming a typical wash-out suppression of order $10^{-2}$, the observed baryon asymmetry can be obtained for values of $(v,v')$ close to the range selected to fit charged lepton masses and mixing angles. Another class of models where the $C\!P$ asymmetries vanish is the one of type I see-saw models where the Dirac and Majorana neutrino mass matrix $m_\nu^D$ and $M$ as well as their see-saw combination $m_\nu$ are form-diagonalizable. A matrix $A$ depending on a set of parameters $\alpha_i$ is said to be form-diagonalizable \cite{Low:2003} if it is diagonalized by unitary transformations that do not depend on $\alpha_i$: \begin{equation} U_L^\dagger A(\alpha) U_R = A^d(\alpha) \end{equation} where $A^d(\alpha)$ is diagonal and the unitary matrices $U_{L,R}$ are independent from $\alpha$. Examples of form-diagonalizable matrix are $m_\nu$ in eqs. (\ref{1k1},\ref{4k1}) and (\ref{1k},\ref{4k}). The parameters are the eigenvalues $m_{1,2,3}$ and the diagonalizing matrices are $U_R=U_L^*=U_{TB}$ and $U_R=U_L^*=U_{BM}$, respectively. As we have seen in section 2, form-diagonalizable matrices naturally arise in the context of models with discrete flavor symmetries. It is possible to show that if in a type I see-saw $m_\nu^D$, $M$ and $m_\nu$ are all form-diagonalizable, then the matrix $\hat{\cal Y}$ is diagonal and the $C\!P$ asymmetries vanish \cite{Aristizabal:2009,Felipe:2009}. In realistic models $m_\nu^D$, $M$ and $m_\nu$ are typically form-diagonalizable only in some symmetry limit. Symmetry breaking terms usually spoil this property and allow for small non-vanishing $C\!P$ asymmetries. So far we have discussed the regime of large RH masses and large neutrino Yukawa couplings. When the smallest RH neutrino mass is below the so-called Davidson-Ibarra bound \cite{Davidson:2002} $(4\times 10^8\div 2\times 10^{9})$ GeV, and we are in the regime of strong hierarchy among RH neutrino masses, the $C\!P$ asymmetry associated to the lightest RH neutrino decay is too small to allow for a successful leptogenesis. To evade the Davidson-Ibarra bound we should depart from the strong hierarchical regime. Under certain conditions a significant enhancement of the $C\!P$ asymmetry can be achieved even for RH neutrino mass ratios as small as 0.1 \cite{Hambye:2004,Raidal:2006}. Alternatively, we can exploit the regime of resonant leptogenesis \cite{Pilaftsis:1997}, occurring when the decaying RH neutrino is quasi-degenerate in mass with some other RH neutrino, the mass differences being comparable with the RH neutrino decay width. A quasi-degeneracy of the RH neutrino spectrum is better understood and dynamically controlled in the presence of an underlying flavor symmetry. Several symmetries have been proposed in the literature such as $G_f=SU(3)_{e^c}\times SU(3)_l\times O(3)_{\nu^c}$ in minimal lepton flavor violation \cite{Cirigliano:2005,Cirigliano:2006,Davidson:2006,Cirigliano:2007,Cirigliano:2007a,Branco:2007} or $G_f=SO(3)$ in ref. \cite{Pilaftsis:2005}. In these two examples the light neutrino masses and their mixing angles are not explained but just accommodated. An interesting model based on a flavor symmetry group $G_f=A_4\times Z_3\times Z_4$ is that of ref. \cite{Branco:2009}. Like the model discussed in Sect. 4 it predicts a lepton mixing close to TB. Due to the presence of an additional discrete factor in the symmetry group, the RH neutrino spectrum is degenerate at LO, and the degeneracy is lifted by radiative corrections or small soft breaking terms, allowing for successful resonant leptogenesis, for a wide range of RH neutrino masses. \section{Summary and conclusion} We have reviewed the motivation, the formalism and the implications of applying non abelian discrete flavor groups to the theory of neutrino mixing. The data on neutrino mixing are by now quite precise. It is a fact that, to a precision comparable with the measurement accuracy, the TB mixing pattern is well approximated by the data (see Fig. (2)). If this experimental result is not a mere accident but a real indication that a dynamical mechanism is at work to guarantee the validity of TB mixing in the leading approximation, corrected by small non leading terms, then non abelian discrete flavor groups emerge as the main road to an understanding of this mixing pattern. Indeed the entries of the TB mixing matrix are clearly suggestive of "rotations" by simple, very specific angles. In fact the group $A_4$, the simplest group used to explain TB mixing, is specified by the set of those rotations that leave a regular tetrahedron invariant. We have started by recalling some basic notions about finite groups and then we have concentrated on those symmetries, like $A_4$ and $S_4$, that are found to be the main candidates for obtaining TB mixing. We have discussed the general mechanism that realizes TB mixing within the framework of discrete flavor symmetries. The symmetry is broken down to two different subgroups in the charged lepton sector and in the neutrino sector, and the mixing matrix arises from the mismatch between the two different residual simmetries. TB mixing requires a flavor symmetry group possessing appropriate residual subgroups. The breaking can be realized in a natural way through the specific vacuum alignments of a set of scalar flavons. We have described a set of models where TB mixing is indeed derived at leading order within this mechanism. There are many variants of such models (in particular with or without see-saw) with different detailed predictions for the spectrum of neutrino masses and for deviations from the TB values of the mixing angles. In general at NLO the different mixing angles receive corrections of the same order of magnitude, which are constrained to be small due to the experimental results which are very close to the TB values. Indeed the small experimental error on $\theta_{12}$, which nicely agrees with the value predicted by TB mixing, suggests that the NLO corrections should be of order of few percent, at most. Additional symmetries are needed, typically of the $U(1)_{FN}$ or $Z_N$ type, in order to reproduce the mass hierarchy of charged leptons. In the neutrino sector there is no reason for the mass eigenvalues not to be of the same order in absolute value. Thus the smallness of the ratio $\sqrt{r} \sim 0.2$, where $r$ is defined in eq. (\ref{r}), is accidental in most of these models. Both normal and inverse hierarchy spectra can be realized. The phenomenology of the models was summarized. We have also discussed the implications of models based on discrete flavor groups for lepton flavor violation and for leptogenesis. Lepton flavor violating processes, the muon g-2 and the EDM's of leptons impose strong constraints on every new physics model. This is also true for the models considered here. But the specific suppression factors and selection rules induced by the finite flavor symmetry group, in particular by $A_4$, may help to improve the consistency of the model even in the presence of new physics at the TeV scale. The observed baryon asymmetry in the Universe, explained in terms of leptogenesis from the decay of heavy Majorana neutrinos, is found to be compatible with models based on discrete groups. Neutrino Yukawa couplings of order one and RH neutrino masses of order $10^{14}\div 10^{15}$ GeV would typically lead to CP asymmetries too large to reproduce the observed baryon asymmetry. However, as a consequence of a general group theoretical property, in all models where the three RH neutrinos transform in a single irreducible representation of the flavor group, the unflavored CP asymmetries vanish in the limit of exact symmetry and small values can be generated through NLO corrections. An obvious question is whether some additional indication for discrete flavor groups can be obtained by considering the extension of the models to the quark sector, perhaps in a Grand Unified context. The answer appears to be that, while the quark masses and mixings can indeed be reproduced in models where TB mixing is realized in the leptonic sector through the action of discrete groups, there are no specific additional hints in favour of discrete groups that come from the quark sector. Examples of Grand Unified descriptions of all fermion masses and mixings with TB mixing for neutrinos have been produced and have been discussed in this review. For quarks, only the third generation masses are present at leading order in these models. The other entries of the mass matrices are small due to additional symmetries or other dynamical reasons (for example, suppression factors from extra dimensions), and the small mass ratios and the small mixing angles are generated by these corrective effects and are not due to the discrete group. As a consequence, the action of the discrete flavor group is only clearly manifest among the comparable neutrino sector masses, in the basis where charged leptons are diagonal. Different forms of neutrino mixing other than TB mixing are also amenable to a description in terms of discrete groups. In alternative to TB mixing, in sect. 8 we have discussed the possibility that actually a more appropriate starting point, could be BM mixing, corrected by large terms of $O(\lambda_C)$, with $\lambda_C$ being the Cabibbo angle ("weak complementarity"), arising from the diagonalization of charged leptons. By suitably modifying the construction in terms of discrete groups adopted in the case of TB mixing, we have identified the group $S_4$ as a good candidate to also provide, in a different presentation, the basis for naturally obtaining BM mixing in first approximation. In the model described the NLO terms are such that the dominant corrections only affect $\theta_{12}$ and $\theta_{13}$ (which receive $O(\lambda_C)$ shifts), while $\theta_{23}$ receives smaller corrections. A value of $\theta_{13}$ near the present bound would support this possibility. In the near future the improved experimental precision on neutrino mixing angles, in particular on $\theta_{13}$, could make the case for TB mixing stronger and then, as a consequence, also the case for discrete flavor groups would be strenghtened. Further important input could come from the LHC. In fact, new physics at the weak scale could have important feedback on the physics of neutrino masses and mixing. \section*{Acknowledgements} We recognize that this work has been partly supported by the Italian Ministero dell'Universit\`a e della Ricerca Scientifica, under the COFIN program (PRIN 2008) and by the European Commission under the networks "Heptools" and "Quest for Unification" and contracts, MRTN-CT-2006-035505 and PITN-GA-2009-237920 (UNILHC). \vfill \newpage \providecommand{\newblock}{}
\section{Introduction Nonlinear Hamiltonian lattices like chains of interacting atoms or coupled oscillators are ubiquitous in mathematics, physics, and material sciences. The most famous example, and most elementary model for a crystal, is a chain of identical atoms that interact by nearest neighbour forces. In reminiscence of the pioneering paper by Fermi, Pasta, and Ulam \cite{FPU55} one usually refers to such systems as FPU or FPU-type chains. \par Although FPU chains are quite simple lattice models they exhibit a rich and complicate dynamical behaviour, and we still lack a complete understanding of their dynamical properties. A mayor topic in the analysis of FPU chains is therefore the investigation of coherent structures such as travelling waves and breathers. Travelling waves are highly symmetric, exact solutions to the underlying lattice equation. They can be regarded as the fundamental modes of nonlinear wave propagation and provide much insight into the energy transport in discrete media. In this paper we aim in contributing to the general theory by studying fronts, i.e., heteroclinic travelling waves that connect two different constant states. \par\quad\newline\noindent The dynamics of FPU chains is governed by the lattice equation \begin{align}% \label{Eqn:Intro.FPU1} \ddot{x}_j=\Phi^\prime\at{x_{j+1}-x_{j}}-\Phi^\prime\at{x_{j}-x_{j-1}}. \end{align} Here $x_j=x_j\at{t}$ denotes the position of the $j^{th}$ atom at time $t$, $\Phi$ is the interaction potential, and the atomic mass is normalized to $1$. Introducing the atomic distances $r_j=x_{j+1}-x_{j}$ and velocities $v_{j}=\dot{x}_{j}$ we can reformulate \eqref{Eqn:Intro.FPU1} as \begin{align}% \label{e:FPU} \dot{r}_j = v_{j+1}-v_j\;,\qquad \dot{v}_j = \Phi^{\prime}\at{r_j} - \Phi^{\prime}\at{r_{j-1}}. \end{align} A travelling wave is a special solution to \eqref{e:FPU} that satisfies the ansatz \begin{align} \label{e:TW.Ansatz} r_j\att=R\at{j-{\sigma}{t}},\qquad v_j\att=V\at{j-{\sigma}{t}}. \end{align} $R$ and $V$ are the \emph{profile functions} for distances and velocities, ${\sigma}$ denotes the wave speed, and ${\varphi}=j-{\sigma}{t}$ is the \emph{phase variable}. In dependence of the properties of $R$ and $V$ travelling waves come in different types. \emph{Wave trains} have periodic profiles and are investigated in \cite{FV99,PP00,DHM06}. They describe oscillatory solutions to \eqref{Eqn:Intro.FPU1} and provide the building blocks for Whitham's modulation theory. Another important class of travelling waves are \emph{solitons} (or \emph{solitary waves}), where $R$ and $V$ are localized over a constant background state. The existence of solitons in lattices is a nontrivial problem and has been studied intensively during the last 20 years. We refer to \cite{FW94,SW97,FM02,Pan05,SZ07,Her10a} for variational methods, and to \cite{Ioo00,IJ05} for an approach via spatial dynamics and centre manifold reduction. \par\quad\newline\noindent In this paper we study \emph{fronts} which have heteroclinic shape and satisfy \begin{align}% \label{e:AsymptoticStates} % \lim\limits_{{\varphi}\to\pm\infty}R\at{\varphi}=r_\pm ,\qquad% \lim\limits_{{\varphi}\to\pm\infty}V\at{\varphi}=v_\pm \end{align}% with $\pair{r_-}{v_-}\neq\pair{r_+}{v_+}$. Fronts have attracted much less interest than solitons, maybe because they only exist if the asymptotic states satisfy some very restrictive conditions. In particular, $\Phi^\prime$ must have at least one turning point between $r_-$ and $r_+$, and this excludes for instance the famous Toda potential. Nonetheless, fronts in FPU chains appear naturally in atomistic Riemann problems, see \cite{HR10a} for numerical simulations, and are important in the context of phase transitions. \par The first rigorous result about fronts we are aware of is the bifurcation criterion from \cite{Ioo00}. It implies that fronts with small jumps between the asymptotic states exist only if $\Phi^\prime$ has a convex-concave turning point. Recently, the existence of fronts was proven by variational methods in \cite{HR10b}. The existence theorem therein does not require the asymptotic states to be close to each other but is restricted to convex potentials $\Phi$. The proof relies on a Lagrangian action integral for fronts with prescribed asymptotic states and uses the direct approach to establish the existence of minimizers. A similar approach is used in \cite{KZ09a,KZ09b} to prove the existence of fronts for sine-Gordon chains. \par In this paper we generalize the method from \cite{HR10b} and prove the existence of fronts without convexity assumption on $\Phi$. Our main result can be summarized as follows. \begin{theorem} \label{Intro:MainTheo} Action minimizing front solutions to \eqref{e:FPU} exist under the following hypotheses: \begin{enumerate} \item[$\at{i}$] The asymptotic states and the front speed satisfy the \emph{macroscopic constraints}, which take the form of three independent jump conditions. \item[$\at{ii}$] The potential satisfies the \emph{graph condition} with respect to the asymptotic states. \item[$\at{iii}$] Some technical assumptions are also satisfied. \end{enumerate} Moreover, there is no front without $\at{i}$, and no action minimizing front without $\at{ii}$. \end{theorem} The assumptions in Theorem \ref{Intro:MainTheo} will be specified below. The macroscopic constraints, see Lemma \ref{Lem:JumpCond}, are algebraic relations and link fronts to energy conserving shocks of the p-system, which is the na\"{\i}ve continuum limit of FPU chains. In particular, they determine the wave speed ${\sigma}$ and imply that the asymptotic strains $r_-$ and $r_+$ cannot be chosen independently of each other. The graph condition reformulates the area condition from \cite{HR10b} and requires that the graph of $\Phi$ is below the shock parabola associated with the asymptotic states. Both the macroscopic constraints and the graph condition appear naturally in our variational existence proof and guarantee that the action integral is well-defined and bounded from below. \par\quad\newline\noindent Closely related to fronts are heteroclinic waves with oscillatory tails. These are travelling wave solutions to \eqref{e:FPU} which approach two different periodic waves for ${\varphi}\to\pm\infty$. Such oscillatory fronts are used to describe martensitic phase transitions and to derive kinetic relations in solids \cite{BCS01a,BCS01b,AP07,Vai10}. The only available existence results, however, concern piecewise quadratic potentials, which allow for simplifying the travelling wave equation by means of Fourier transform, see \cite{TV05,SCC05,SZ09}. It remains a challenging problem for future research to give alternative, maybe variational, existence proofs that cover more general chains. \par\quad\newline\noindent The paper is organized as follows. In \S\ref{sec:waves} we discuss the macroscopic constraints and normalize the asymptotic states. Moreover, we reformulate the front equation as an eigenvalue problem for a nonlinear integral operator. In \S\ref{sec:proof} we set the existence problem into a variational framework and characterize fronts as minimizers of an action integral. Or main technical result is Theorem \ref{Theo:Minimiser} and guarantees that this action integral attains its minimum on a suitable set of candidates for fronts. The proof uses \emph{separations of phases}, which are introduced in \S\ref{sec:proof:sop} and allow to extract convergent subsequences from action minimizing sequences. Finally, we present some numerical simulations in \S\ref{sec:num}. \section{Preliminaries about fronts}\label{sec:waves} Substituting the travelling wave ansatz \eqref{e:TW.Ansatz} into \eqref{e:FPU} yields \begin{align}% \label{e:tw} % {\sigma}\tfrac{\dint}{\dint{\varphi}}R({\varphi})+V({\varphi}+1)-V({\varphi})=0,\qquad {\sigma}\tfrac{\dint}{\dint{\varphi}}V({\varphi})+ \Phi^\prime\bat{R\at{{\varphi}}}- \Phi^\prime\bat{R\at{{\varphi}-1}}=0, \end{align} which is a nonlinear system of advance-delay-differential equations. Moreover, combining both equations we readily verify the energy law \begin{align}% \label{e:tw.energy} % {\sigma}\tfrac{\dint}{\dint{\varphi}}\Bat{\tfrac{1}{2}V^2\at{\varphi}+\Phi\at{R\at{\varphi}}}+ \Phi^\prime\bat{R\at{\varphi}}V\at{{\varphi}+1}- \Phi^\prime\bat{R\at{{\varphi}-1}}V\at{{\varphi}}=0. \end{align} \subsection{Macroscopic constraints for the asymptotic states}% We now derive the macroscopic constraints that couple the front speed ${\sigma}$ to the asymptotic states $\pair{r_\pm}{v_\pm}$ from \eqref{e:AsymptoticStates}. To this end we consider continuous observables $\psi=\psi\pair{r}{v}$ and denote by \begin{align*} \jump{\psi\pair{r}{v}}:=\psi\pair{r_+}{v_+}-\psi\pair{r_-}{v_-} \qquad\text{and}\qquad \mean{\psi\pair{r}{v}}:=\tfrac{1}{2}\at{\psi\pair{r_-}{v_-}+\psi\pair{r_+}{v_+}}, \end{align*} the \emph{jump} and \emph{mean value}, respectively. \par The following result was proven in \cite{HR10b} (see also \cite{AP07}) by integrating \eqref{e:tw} and \eqref{e:tw.energy} over a finite interval $\ccinterval{-N}{N}$ and passing to the limit $N\to\infty$. \begin{lemma}% \label{Lem:JumpCond} The asymptotic states of each front satisfy \begin{align} \label{Lem:JumpCond.Eqn1} {\sigma}\jump{r}+\jump{v}=0,\qquad {\sigma}\jump{v}+\jump{\Phi^\prime\at{r}}=0,\qquad {\sigma}\jump{\tfrac{1}{2}v^2+\Phi\at{r}}+\jump{\Phi^\prime\at{r}v}=0. \end{align} \end{lemma}% Heuristically, Lemma \ref{Lem:JumpCond} reflects that fronts transform into shock waves when passing to large spatial and temporal scales. The jump conditions \eqref{Lem:JumpCond.Eqn1} precisely mean that the asymptotic states correspond to an \emph{energy conserving shock} for the p-system and imply that each front satisfies mass, momentum, and energy. The p-system is the na\"{\i}ve continuum limit of FPU chains under the hyperbolic scaling and reads \begin{align} \label{Eqn:PSystem} \partial_\tau{r}=\partial_y{v},\qquad \partial_\tau{v}=\partial_y{\Phi^\prime\at{r}}, \end{align} where $\tau={\varepsilon}{t}$ and $y={\varepsilon}{j}$ denote the macroscopic time and space, respectively, and ${\varepsilon}>0$ is a small scaling parameter. The conservation laws in \eqref{Eqn:PSystem} correspond to mass and momentum, and imply the conservation of energy for smooth solutions, that is \begin{align} \label{Eqn:PSystem.E} \partial_\tau\at{\tfrac{1}{2}v^2+\Phi\at{r}}=\partial_\tau\at{v\Phi^\prime\at{r}}. \end{align}% The jump conditions for \eqref{Eqn:PSystem.E}, however, is independent of the jump conditions for \eqref{Eqn:PSystem}. More details about the p-system and energy conserving shocks can be found in \cite{HR10a,HR10b}. \par Using the discrete Leibniz rule % \begin{math} % \jump{\psi_1 \psi_2}=\jump{\psi_1}\mean{\psi_2}+\mean{\psi_1}\jump{\psi_2} \end{math} % we readily verify that \eqref{Lem:JumpCond.Eqn1} implies \begin{align} \label{Lem:JumpCond.Eqn2} \jump{\Phi\at{r}}=\jump{r}\mean{\Phi^\prime\at{r}} ,\qquad% {\sigma}^2=\jump{\Phi^\prime\at{r}}/\jump{r}. \end{align} Conversely, for any $\pair{r_-}{r_+}$ with \eqref{Lem:JumpCond.Eqn2}$_1$ there exist -- up to Galilean transformations -- exactly two solutions to \eqref{Lem:JumpCond.Eqn1} which differ in $\mathrm{sgn}{{\sigma}}$. We now characterize the geometric meaning of \eqref{Lem:JumpCond.Eqn2} and refer to Figure \ref{fig:shocks} for an illustration. \begin{figure}[ht!] \centering{% \includegraphics[width=0.975\textwidth, draft=\figdraft]% {\figfile{shocks}}% \caption{% To each front there exists a parabola that touches the graph of $\Phi$ in both $r_-$ and $r_+$. Consequently, the signed area between the graph of $\Phi^\prime$ and the secant connecting $r_-$ and $r_+$ vanishes the stripe $\ccinterval{r_-}{r_+}$. }% \label{fig:shocks} }% \end{figure} \begin{lemma} The following conditions are equivalent: \begin{enumerate} \item[$\at{i}$] $\triple{{\sigma}}{r_-}{r_+}$ fulfils \eqref{Lem:JumpCond.Eqn2}, \item[$\at{ii}$] there exists a parabola that touches the graph of $\Phi$ in both $r_-$ and $r_+$, \item[$\at{iii}$] the signed area between the graph of $\Phi^\prime$ and the secant connecting $r_-$ to $r_+$ sums up to zero in $\ccinterval{r_-}{r_+}$. \end{enumerate} Moreover, each condition implies that $\Phi^\prime$ has at least one turning point between $r_-$ and $r_+$. \end{lemma} \begin{proof} Consider the parabola $f\at{r}=\tfrac{1}{2}ar^2+br+c$. The touching conditions \begin{align*} f\at{r_\pm}=\Phi\at{r_\pm},\quad f^\prime\at{r_\pm}=\Phi^\prime\at{r_\pm} \end{align*} are equivalent to \begin{align*} \tfrac{1}{2}a\jump{r^2}+b\jump{r}=\jump{\Phi\at{r}},\quad \tfrac{1}{2}a\mean{r^2}+b\mean{r}+c=\mean{\Phi\at{r}},\quad \jump{r}a=\jump{\Phi^\prime\at{r}} ,\quad a\mean{r}+b=\mean{\Phi^\prime\at{r}}, \end{align*} and by $\tfrac{1}{2}\jump{r^2}=\mean{r}\jump{r}$ we conclude that $\at{i}$ and $\at{ii}$ are equivalent via \begin{align*} a={\sigma}^2 ,\quad b=\mean{\Phi^\prime\at{r}}-{\sigma}^2\mean{r} ,\quad {c}=& \mean{\Phi\at{r}}-\mean{r}\mean{\Phi^\prime\at{r}}+ {\sigma}^2\bat{\mean{r}^2-\tfrac{1}{2}\mean{r^2}}. \end{align*} The equivalence of $\at{ii}$ and $\at{iii}$ is immediate since the secant has slope ${\sigma}^2$, and $\Phi^\prime$ must have a turning point because otherwise the graph of $\Phi^\prime$ would be either below or above the secant. \end{proof} Condition \eqref{Lem:JumpCond.Eqn2}$_1$ is the \emph{kinetic relation} for fronts and reveals that the asymptotic states cannot be chosen arbitrarily. More precisely, for given $r_-$ and $r_+$ we can choose ${\sigma}$ and $\jump{v}$ such that the first two jump conditions in \eqref{Lem:JumpCond.Eqn1} (which correspond to mass and momentum) are satisfied. However, for the energy condition \eqref{Lem:JumpCond.Eqn1}$_3$ to hold, $r_-$ and $r_+$ must additionally fulfil \eqref{Lem:JumpCond.Eqn2}$_1$. Form this we conclude that fronts do not exist if $\Phi^\prime$ is either convex or concave, and that in general we cannot prescribe both $r_-$ and $r_+$. \par\quad\newline\noindent We emphasize that \eqref{Lem:JumpCond.Eqn1} is in general not sufficient for the existence of fronts, {i.e.}, there exist energy conserving shocks in the p-system that can not be realized by a front in FPU. In fact, it was proven in \cite{Ioo00} that fronts bifurcate from convex-concave but not from concave-convex turning points of $\Phi^\prime$. This disproves the existence of \emph{subsonic} fronts with small jump heights although there exist the corresponding energy conserving shocks. \par In order to prove the existence of action minimizing fronts we shall additionally to \eqref{Lem:JumpCond.Eqn1} require that the graph of $\Phi$ is below the parabola defined by the asymptotic states, see Assumption \ref{Ass:Pot}. In particular, our existence result provides a front for Example $A$ from Figure \ref{fig:shocks} but does not cover Example $B$, see Remark \ref{Rem:UnboundedL} and the examples in \S\ref{sec:num}. \subsection{Normalization and reformulation}% For our analysis in \S\ref{sec:proof} it is convenient to normalize the asymptotic states and to reformulate the front equation \eqref{e:tw} as an eigenvalue problem for a nonlinear integral operator. \begin{lemma} \label{Lem:Normalisation} Up to affine transformations we can assume that \begin{align} \label{Norm.Problem.States} {\sigma}=1 ,\quad% r_\pm=\pm1 ,\quad% v_\pm=\mp1 ,\quad% \Phi^\prime\at{\pm1}=\pm1 ,\quad% \Phi\at{\pm1}=\tfrac{1}{2}. \end{align} Moreover, with \eqref{Norm.Problem.States} the front equation is equivalent to \begin{align}% \label{Norm.Problem.Eqn}% W=\mathcal{A}{\Phi}^\prime\at{\mathcal{A}{W}} ,\qquad% \at{\mathcal{A}{W}}\at{\varphi}= \int\limits_{{\varphi}-\tfrac{1}{2}}^{{\varphi}-\tfrac{1}{2}}W\at{\tilde{\varphi}}\dint\tilde{\varphi}, \end{align} where $W$ is a normalized profile with $\lim_{{\varphi}\to\pm\infty}W\at{\varphi}=\pm1$. \end{lemma} \begin{proof} Let $U$ and $W$ be two normalized profiles such that \begin{align*} {R}\at{\varphi}=\mean{r}+\tfrac{1}{2}\jump{r}U\at{{\varphi}+1/2} ,\qquad% V\at{\varphi}=\mean{v}+\tfrac{1}{2}\jump{v}W\at{\varphi}. \end{align*} Using the first two jump conditions from \eqref{Lem:JumpCond.Eqn1} we readily verify that \eqref{e:tw} transforms into \begin{align}% \label{Lem:Normalisation.Eqn1}% \tfrac{\dint}{\dint{\varphi}}U\at{\varphi}=W\at{{\varphi}+1/2}-W\at{{\varphi}-1/2} ,\qquad% \tfrac{\dint}{\dint{\varphi}}W\at{\varphi}=\wh{\Phi}^\prime\bat{U\at{{\varphi}+1/2}}- \wh{\Phi}^\prime\bat{U\at{{\varphi}-1/2}}, \end{align} where the normalized potential \begin{align*} \wh{\Phi}\at{u}=\frac{4}{\jump{\Phi^\prime\at{r}}\jump{r}} {\Phi}\Bat{\mean{r}+\tfrac{1}{2}\jump{r}\,u} -\frac{2\mean{\Phi^\prime\at{r}}}{\jump{\Phi^\prime\at{r}}}\,{u}+\frac{1}{2}- \frac{4\mean{\Phi\at{r}}}{\jump{\Phi^\prime\at{r}}\jump{r}} \end{align*} satisfies $\wh{\Phi}^\prime\at{\pm1}=\pm1$. Moreover, we have $\wh{\Phi}\at{-1}=\wh{\Phi}\at{+1}=\tfrac{1}{2}$ if and only if the third jump condition \eqref{Lem:JumpCond.Eqn1}$_3$ is satisfied. Towards \eqref{Norm.Problem.Eqn} now suppose \eqref{Norm.Problem.States}. Integrating \eqref{Lem:Normalisation.Eqn1}$_1$ we find $U=\mathcal{A}{W}$, where the constant of integration vanishes due to $U\at{\pm1}=W\at{\pm1}=\pm1$, and similarly we derive $W=\mathcal{A}\wh{\Phi}^\prime\at{U}$ from \eqref{Lem:Normalisation.Eqn1}$_2$. \end{proof} The front parabola for normalized data \eqref{Norm.Problem.States} is $r\mapsto\tfrac{1}{2}r^2$ and each solution to \eqref{Norm.Problem.Eqn} can be viewed as a perturbation of the \emph{shock profile} \begin{align} \label{E:DefShock}% W_{\rm sh}\at{\varphi}=\mathrm{sgn}{\varphi}=\left\{ \begin{array}{rcl} +1&\text{for}&{\varphi}<0,\\ 0&\text{for}&{\varphi}=0,\\ -1&\text{for}&{\varphi}>0. \end{array} \right. \end{align} Notice that the residual of $W_{\rm sh}$, that is $W_{\rm sh}-\mathcal{A}\Phi^\prime\at{\mathcal{A}{W_{\rm sh}}}$, has compact support. \par\quad\newline\noindent We proceed with some preliminary remarks about the action of a front. Heuristically, the action density in the normalized setting is given by \begin{align*} \tfrac{1}{2}W^2-\Phi\at{\mathcal{A}{W}} =\tfrac{1}{2}W^2-\tfrac{1}{2}\at{\mathcal{A}{W}}^2+\Psi\at{\mathcal{A}{W}} \end{align*} with \begin{align} \label{E:DefPsi}% \Psi\at{r}=\tfrac{1}{2}r^2-\Phi\at{r}, \end{align} so the action integral formally reads \begin{align} \label{E:DefAction2}% \wt{\mathcal{L}}\at{W}&= \int\limits_{\mathbb{R}} \tfrac{1}{2}W^2-\tfrac{1}{2}\at{\mathcal{A}{W}}^2+\Psi\at{\mathcal{A}{W}} \dint{\varphi}. \end{align} Notice that $\Psi$ is just the difference between the front parabola and $\Phi$, see Figure \ref{fig:shocks}, and that $\wt{\mathcal{L}}$ is well defined as long as $W$ approaches its asymptotic states sufficiently fast. A further possibility for defining the action integral was introduced in \cite{HR10b} for monotone $W$ and relies on the relative action integral \begin{align*} \wh{\mathcal{L}}\at{W}&= \int\limits_{\mathbb{R}} % \Bat{\tfrac{1}{2}W^2- \Phi\at{\mathcal{A}{W}}}- \Bat{ \tfrac{1}{2}W_{\rm sh}^2- \Phi\at{\mathcal{A}{W_{\rm sh}}}} \dint{\varphi}. \end{align*} Both approaches are linked by $\wh{\mathcal{L}}\at{W}=\wt{\mathcal{L}}\at{W}-\wt{\mathcal{L}}\at{W_{\rm sh}}$ and the symmetry of $\mathcal{A}$, compare Lemma \ref{Lem:AProps}, formally implies \begin{align*} \partial\wh{\mathcal{L}}\at{W}=\partial\wt{\mathcal{L}}\at{W}=W-\mathcal{A}\Phi^\prime\at{\mathcal{A}{W}}. \end{align*} In \S\ref{sec:proof} we give a slightly different definition of $\wt\mathcal{L}$, see \eqref{Eqn:DefM} and \eqref{Eqn:DefL}, and establish the existence of minimizers. \section{Existence of fronts}\label{sec:proof} In this section we assume that the asymptotic states and the potential are normalized by \eqref{Norm.Problem.States} and show that the fixed point equation \eqref{Norm.Problem.Eqn} has a solution in some appropriate function space. \subsection{Assumptions}% \newcommand{\cond}[1]{(#1)} We rely on the following standing assumptions on the function $\Psi$ from \eqref{E:DefPsi}. Examples and counterexamples are given in \S\ref{sec:num}. \begin{assumption} \label{Ass:Pot}% $\Psi$ is continuously differentiable and satisfies the following conditions: \begin{enumerate} \item[\cond{G}] \emph{graph condition}: $\Psi\at{u}\geq\Psi\at{\pm1}=0$ for all $u\in{\mathbb{R}}$, \item[\cond{X}] \emph{genericity}: $\Psi^{\prime\prime}\at{\pm1}>0$ and $\Psi\at{u}>0$ for all $u\neq\pm1$, \item[\cond{M}] \emph{monotone asymptotic behaviour}: $\Psi\at{u}$ is decreasing for $u\ll-1$ and increasing for $u\gg1$. \end{enumerate}% \end{assumption} Condition \cond{G} has a natural interpretation in terms of $\Phi$ and can easily be reformulated for non-normalized data: It precisely means that the front parabola touches the graph of $\Phi$ in both $r_-$ and $r_+$ but is above this graph in all other points. Moreover, \cond{G} is equivalent to the \emph{area condition} from \cite{HR10b}, which characterizes the signed area between the graph of $\Phi^\prime$ and the secant connecting $r_-$ to $r_+$. With positive and negative sign for above and below the graph of $\Phi^\prime$, respectively, the area condition reads as follows. The signed area is non-negative in each stripe $\ccinterval{r_-}{r}$ with $r>r_-$ and non-positive in each stripe $\ccinterval{r}{r_+}$ with $r<r_+$. We refer to Figure \ref{fig:shocks} for illustration, where positive and negative area are displayed in dark and light grey colour, respectively, and recall that the signed area vanishes in the stripe $\ccinterval{r_-}{r_+}$ due to \eqref{Lem:JumpCond.Eqn2}$_1$. \par We mention that \cond{G} is truly necessary for the existence of action minimizing fronts, see Remark \ref{Rem:UnboundedL}. The conditions \cond{M} and \cond{X}, however, are made for convenience and might be weakened for the price of more technical effort. \begin{remark} % \cond{X} is equivalent to \begin{enumerate} \item[\cond{S}] \emph{supersonic front speed}: $\Phi^{\prime\prime}\at{\pm1}<1$, \end{enumerate} and \cond{M} implies \begin{enumerate} \item[\cond{I}] \emph{invariant set for $\Phi^{\prime}$}: There exits a constant $\Gamma>1$ such that $\Phi^\prime$ maps $\ccinterval{-\Gamma}{\Gamma}$ into itself. \end{enumerate} \end{remark} \begin{proof} \cond{S} follows from the definition of $\Psi$ in \eqref{E:DefPsi}. Towards \cond{I} we exploit \cond{M} to choose $\tilde{\Gamma}>1$ such that $\Phi^\prime\at{u}>u$ for $u<-\tilde{\Gamma}$ and $\Phi^\prime\at{u}<u$ for $u>\tilde{\Gamma}$. Then we set \begin{math} \Gamma=\max\{\tilde{\Gamma},\,\max_{\abs{u} \leq% {\tilde{\Gamma}}}\abs{\Phi^{\prime}\at{u}}\} \end{math}. \end{proof} \subsection{Functionals and operators}% We denote by $\fspace{L}^p$, $\fspace{W}^{1,\,p}$ and $\fspace{C}^k$ the usual function spaces on the real line, abbreviate the $\fspace{L}^p$-norm by $\norm{\cdot}_p$, and write \begin{align*} \skp{W_1}{W_2}=\int_{\mathbb{R}}{W_1\at{\varphi}}{W_2}\at{\varphi}\dint{\varphi} \end{align*} for the dual pairing of $W_1\in\fspace{L}^p$ and $W_2\in\fspace{L}^{p'}$ with $1=1/p+1/p'$. \begin{lemma} \label{Lem:AProps} The averaging operator $\mathcal{A}$ has the following properties: \begin{enumerate} \item % $\mathcal{A}$ maps $\fspace{L}^p$ into $\fspace{L}^\infty\cap\fspace{W}^{1,\,p}\subset\fspace{C}$ for all $1\leq{p}\leq\infty$ with \begin{align*} \at{\mathcal{A}{W}}^\prime\at{\varphi} =% W\at{{\varphi}+\tfrac{1}{2}}-W\at{{\varphi}-\tfrac{1}{2}} \end{align*} % and $\norm{\mathcal{A}{W}}_{\fspace{L}^p}\leq\norm{W}_{\fspace{L}^p}$, $\norm{\mathcal{A}{W}}_{\fspace{L}^\infty}\leq\norm{W}_{\fspace{L}^p}$, $\norm{\mathcal{A}{W}}_{\fspace{W}^{1,\,p}}\leq3\norm{W}_{\fspace{L}^p}$. \item $\mathcal{A}$ is symmetric in the sense that $\skp{\mathcal{A}{W_1}}{W_2}=\skp{W_1}{\mathcal{A}{W_2}}$ holds for all $W_1\in\fspace{L}^p$ and $W_2\in\fspace{L}^{p'}$. \item % $\mathcal{A}$ is self-adjoint in $\fspace{L}^2$ with spectrum \begin{math} \mathrm{spec}\mathcal{A}=\{\varrho\at{k}\;:\;k\in{\mathbb{R}}\} \end{math} where $\varrho\at{k}=\tfrac{2}{k}\sin\at{\tfrac{k}{2}}$. \end{enumerate} \end{lemma} \begin{proof} The first two statements are straight forward. The third one follows since $\mathcal{A}$ diagonalizes in Fourier space via $\mathcal{A}\mhexp{\mathtt{i}{k}{\varphi}}=\varrho\at{k}\mhexp{\mathtt{i}{k}{\varphi}}$. \end{proof} We now introduce the affine space \begin{align*} \mathcal{H}=\left\{W\;:\;W-W_{\rm sh}\in\fspace{L}^2\right\}, \end{align*} where $W_{\rm sh}$ is the shock profile from $\eqref{E:DefShock}$. Exploiting Lemma \ref{Lem:AProps}, the Taylor expansion of $\Phi^\prime$ around $\pm1$, and the properties of $W_{\rm sh}$ we then find \begin{align} \label{Rem:HProps.Eqn4} \mathcal{A}{W},\;\Phi^\prime\at{\mathcal{A}{W}},\;\mathcal{A}\Phi^\prime\at{\mathcal{A}{W}}\in\mathcal{H},\qquad W-\mathcal{A}{W},\;W-\mathcal{A}^2{W}\in\fspace{L}^2. \end{align} for all $W\in\mathcal{H}$. In view of the action integral \eqref{E:DefAction2} we also define a functional $\mathcal{M}$ on $\fspace{L}^2$ by \begin{align*} \mathcal{M}\nat{V}= \tfrac{1}{2}\int\limits_{\mathbb{R}}% V^2-\nat{\mathcal{A}{V}}^2 \dint{\varphi}= \tfrac{1}{2}\int\limits_{\mathbb{R}}% \nat{V-\mathcal{A}^2V}V \dint{\varphi}, \end{align*} and a functional $\mathcal{N}$ on $\mathcal{H}$ by \begin{align} \label{Eqn:DefM} \mathcal{N}\at{W}=\mathcal{M}\at{W-W_{\rm sh}}+ \tfrac{1}{2}\int\limits_{\mathbb{R}}% W_{\rm sh}^2-\at{\mathcal{A}{W_{\rm sh}}}^2\dint{\varphi}+ \int\limits_{\mathbb{R}} \at{W-W_{\rm sh}}\at{W_{\rm sh}-\mathcal{A}^2{W_{\rm sh}}}\dint{\varphi}. \end{align} Notice that $\mathcal{N}\at{W}$ is well defined on $\mathcal{H}$ as both $W_{\rm sh}^2-\at{\mathcal{A}{W_{\rm sh}}}^2$ and $W_{\rm sh}-\mathcal{A}^2{W_{\rm sh}}$ have compact support. Moreover, if $W-W_{\rm sh}$ decays sufficiently fast for ${\varphi}\to\pm\infty$ (say $W-W_{\rm sh}\in\fspace{L}^1$), then we have \begin{align} \label{Eqn:MFormula} \mathcal{N}\at{W}= \tfrac{1}{2}\int\limits_{\mathbb{R}}% W^2-\at{\mathcal{A}{W}}^2\dint{\varphi}=\tfrac{1}{2}\int\limits_{\mathbb{R}}% \at{W-\mathcal{A}^2W}W\dint{\varphi}. \end{align} \begin{lemma} \label{Lem:MProps} The functional $\mathcal{M}$ is non-negative and weakly lower semi-continuous on $\fspace{L}^2$. \end{lemma} \begin{proof} Denoting the Fourier transform of $V$ by $\wh{V}$ we find \begin{align} \nota \mathcal{M}\at{V} &=% \int\limits_{\mathbb{R}}\nat{1-\varrho\at{k}^2}\wh{V}\at{k}^2\dint{k} =% \norm{\sqrt{1-\varrho^2}\,\wh{V}}_2 \end{align} with $\varrho$ as in Lemma \ref{Lem:AProps}. This gives the desired result as $V_n\rightharpoonup{V_\infty}$ implies $\wh{V}_n\rightharpoonup{\wh{V}_\infty}$ and hence $\sqrt{1-\varrho^2}\,\wh{V}_n\rightharpoonup{\sqrt{1-\varrho^2}\,\wh{V}_\infty}$ \end{proof} \begin{lemma} \label{Lem:NProps} The functional $\mathcal{N}$ is G\^{a}teaux differentiable on $\mathcal{H}$ with derivative \begin{align} \label{Lem:NProps.Eqn4} \partial\mathcal{N}\at{W}=W-\mathcal{A}^2{W}\in\fspace{L}^2. \end{align} Moreover, $\mathcal{N}$ is invariant under shifts in ${\varphi}$-direction, and satisfies \begin{align} \label{Lem:NProps.Eqn6} \mathcal{N}\at{W_2}=\mathcal{N}\at{W_1}+\mathcal{M}\at{W_2-W_1}+\skp{W_2-W_1}{W_1-\mathcal{A}^2{W}_1} \end{align} for all $W_1,\,W_2\in\mathcal{H}$. \end{lemma} \begin{proof} A direct computation with $W\in\mathcal{H}$ and $\delta{W}\in\fspace{L}^2$ shows \begin{align*} \skp{\partial\mathcal{N}\at{W}}{\delta{W}} &=% \bat{\skp{W-W_{\rm sh}}{\delta{W}} -\skp{\mathcal{A}{W}-\mathcal{A}{W_{\rm sh}}}{\mathcal{A}{\delta{W}}}} +\skp{W_{\rm sh}-\mathcal{A}^2{W_{\rm sh}}}{\delta{W}} \\&=% \skp{W-W_{\rm sh}}{\delta{W}}-\skp{\mathcal{A}^2{W}-\mathcal{A}^2{W_{\rm sh}}}{\delta{W}} +\skp{W_{\rm sh}-\mathcal{A}^2{W_{\rm sh}}}{\delta{W}} \\&=% \skp{W-\mathcal{A}^2}{\delta{W}}, \end{align*} and this gives \eqref{Lem:NProps.Eqn4}. Towards the shift invariance we approximate $W$ by \begin{align*} W_n=\chi_{\ccinterval{-n}{n}}\at{W-W_{\rm sh}}+W_{\rm sh} \end{align*} where $\chi_{\ccinterval{-n}{n}}$ is the indicator function of the interval $\ccinterval{-n}{n}$. Then we use \eqref{Eqn:MFormula} for $W_n$ to find $\mathcal{L}\at{W_n}=\mathcal{L}\at{W_n\at{\cdot+\bar{\varphi}}}$ for all shifts $\bar{\varphi}$, and passing to the limit $n\to\infty$ gives the desired result. Finally, by definition we have \begin{align*} \mathcal{N}\at{W_2}-\mathcal{N}\at{W_1} &=% \mathcal{M}\at{W_2-W_{\rm sh}}-\mathcal{M}\at{W_1-W_{\rm sh}}+\skp{W_2-W_1}{W_{\rm sh}-\mathcal{A}^2W_{\rm sh}} \end{align*} and \begin{align*} \mathcal{M}\at{W_2-W_{\rm sh}}-\mathcal{M}\at{W_1-W_{\rm sh}}= \mathcal{M}\at{W_2-W_1}+ \skp{W_2-W_1}{W_1-W_{\rm sh}-\mathcal{A}^2{W_1}+\mathcal{A}^2{W_{\rm sh}}}, \end{align*} so \eqref{Lem:NProps.Eqn6} follows from adding both identities. \end{proof} To conclude this section we consider the functional \begin{align*} \mathcal{P}\at{W}=% \int\limits_{\mathbb{R}}{}\Psi\at{\mathcal{A}{W}}\dint{\varphi}, \end{align*} which gives the non-quadratic part of the action integral \eqref{E:DefAction2} \begin{lemma} \label{Lem:PProps} $\mathcal{P}$ is well defined on $\mathcal{H}$ with \begin{align} \label{Eqn:LowerEstimateForG} \ul{c}\norm{\mathcal{A}{W}-\mathrm{sgn}\at{\mathcal{A}{W}}}_2 \leq \mathcal{P}\at{W} \leq \ol{c}\norm{\mathcal{A}{W}-\mathrm{sgn}\at{\mathcal{A}{W}}}_2 \end{align} for some constants $\ul{c}$ and $\ol{c}$ that depend only on $\norm{AW}_\infty$. Moreover, $\mathcal{P}$ is G\^{a}teaux differentiable on $\mathcal{H}$ with derivative $\partial\mathcal{P}\at{W}=\mathcal{A}^2{W}-\mathcal{A}\Phi^\prime\at{\mathcal{A}{W}}$. \end{lemma} \begin{proof} Let $W\in\mathcal{H}$ be given and recall that $U=\mathcal{A}{W}$ satisfies $U\in\mathcal{H}\cap\fspace{L}^\infty$ due to Lemma \ref{Lem:AProps}. Condition \cond{X} provides two constants $\ul{c}$ and $\ol{c}$ such that \begin{align*} \ul{c}\at{u-\mathrm{sgn}{u}}^2\leq\Psi\at{u}\leq\ol{c}\at{u-\mathrm{sgn}{u}}^2\quad\text\and\quad \end{align*} holds for all $\abs{u}\leq{\norm{U}_\infty}$, and we conclude that $\mathcal{P}$ is well defined and satisfies \eqref{Eqn:LowerEstimateForG}. Finally, \eqref{Rem:HProps.Eqn4} provides $\mathcal{A}^2{W}-\mathcal{A}\Phi^\prime\at{\mathcal{A}{W}}\in\fspace{L}^2$, so both the existence of and the formula for $\partial\mathcal{P}$ follow from a direct calculation. \end{proof} \subsection{Variational setting}% We now introduce the action functional on $\mathcal{H}$ by \begin{align} \label{Eqn:DefL}% \mathcal{L}\at{W}=\mathcal{N}\at{W}+\mathcal{P}\at{W}. \end{align} In virtue of Lemma \ref{Lem:NProps} and Lemma \ref{Lem:PProps} the functional $\mathcal{L}$ is well defined, shift invariant, and G\^{a}teaux differentiable with derivative \begin{align*} \partial\mathcal{L}\at{W}=\mathcal{A}^2W-\mathcal{A}\Phi^\prime\at{\mathcal{A}{W}}\in\fspace{L}^2, \end{align*} and we conclude that each minimizer of $\mathcal{L}$ in $\mathcal{H}$ must solve the front equation \eqref{Norm.Problem.Eqn}. However, proving the existence of minimizers in $\mathcal{H}$ turns out to be difficult and therefore we restrict $\mathcal{L}$ to the convex subset \begin{align*} \mathcal{C}=\left\{W\in\mathcal{H}\cap\fspace{W}^{1,\infty}% \;:\;% \norm{W}_{\infty}\leq{{\Gamma}} ,\quad% \norm{W^\prime}_{\infty}\leq2{{\Gamma}}\right\}. \end{align*} Notice that the ansatz $W\in\mathcal{C}$ is reasonable due to condition $\cond{I}$ and since the front equation \eqref{Norm.Problem.Eqn} combined with Lemma \ref{Lem:AProps} implies $W\in\mathcal{H}\cap\fspace{W}^{1,\,\infty}$. \par\quad\newline\noindent In order to link fronts to minimizers of $\mathcal{L}$ in $\mathcal{C}$ we observe that the properties of $\Phi^\prime$ and $\mathcal{A}$ guarantee $\mathcal{C}$ to be invariant under the $\fspace{L}^2$-gradient flow of $\mathcal{L}$. To see this we consider the explicit Euler scheme \begin{align} \label{Eqn:EulerScheme} W\mapsto\mathcal{T}_{\lambda}\at{W}=W-{\lambda}\partial\mathcal{L}\at{W}=\at{1-{\lambda}}W+{\lambda}\mathcal{A}\Phi^\prime\at{\mathcal{A}{W}} \end{align} with small step size ${\lambda}$. \begin{lemma} \label{Lem:FrontsAndMinimisers} The set $\mathcal{C}$ is invariant under the action of $\mathcal{T}_{\lambda}$ for $0<{\lambda}<1$. Consequently, each minimizer of $\mathcal{L}$ in $\mathcal{C}$ solves the front equation \eqref{Norm.Problem.Eqn}. \end{lemma} \begin{proof} For $W\in\mathcal{C}$ let $P=\Phi^\prime\at{\mathcal{A}{W}}$ and recall that $\mathcal{A}{W},\;P,\;\mathcal{A}{P}\in\mathcal{H}$ according to \eqref{Rem:HProps.Eqn4}. Combining \cond{I} with $\norm{\mathcal{A}{W}}_\infty\leq{\Gamma}$ and $\at{\mathcal{A}{P}}^\prime=P\at{\cdot+1/2}+P\at{\cdot+1/2}$ gives \begin{align*} \norm{P}_\infty,\;\norm{\mathcal{A}{P}}_\infty\leq{{\Gamma}} ,\qquad\norm{\at{\mathcal{A}{P}}^\prime}_\infty\leq2{{\Gamma}}, \end{align*} and hence $\mathcal{A}{P}\in\mathcal{C}$. Since $\mathcal{C}$ is convex we also have $\mathcal{T}_{\lambda}\at{W}\in\mathcal{C}$ for all $0<{\lambda}<1$, and passing to the limit ${\lambda}\to0$ we then establish the invariance of $\mathcal{C}$ under the $\fspace{L}^2$-gradient flow of $\mathcal{L}$. In particular, each minimizer of $\mathcal{L}$ in $\mathcal{C}$ must be a stationary point for the gradient flow of $\mathcal{L}$ and hence a solution to the front equation. \end{proof} To complete the existence proof for fronts it remains to show that $\mathcal{L}$ attains its minimum in $\mathcal{C}$. We prove this in the next section by using the direct approach, that means we construct minimizers as limits of minimizing sequences. \par A particular problem we have to overcome in the subsequent analysis is that $\mathcal{L}$ is \emph{not} coercive on $\mathcal{H}$. In fact, as illustrated in Figure \ref{fig:sep_phase} there exist sequences $\at{W_n}_n\subset\mathcal{C}$ with extending plateaus at $-1$ or $+1$. These plateaus contribute neither to $\mathcal{N}$ nor $\mathcal{P}$ but may imply \begin{align*} \norm{W_n-W_{\rm sh}\at{{\varphi}_n+\cdot}}_2\xrightarrow{n\to\infty}\infty \end{align*} for all choices of the relative shifts ${\varphi}_n$. Heuristically it is clear that the cartoon from Figure \ref{fig:sep_phase} cannot be prototypical for action minimizing sequences, but in order to proof this we need a better understanding of sequences with bounded action. \begin{figure}[ht!] \centering{% \includegraphics[width=0.7\textwidth, draft=\figdraft]% {\figfile{sep_phase}}% \caption{% Sketch of a sequence with bounded action and two extending plateaus at $\pm1$: graphs of $W_n$ and $U_n=\mathcal{A}{W_n}$ in Black in Gray, respectively. The shaded areas indicate corresponding separations of phases with $I_{n,j}={\varphi}_{n,j}+J_j$ as in Lemma \ref{Lem:SepPhasesForSequ}. }% \label{fig:sep_phase} }% \end{figure} We conclude with a remark about the necessity of the graph condition \cond{G} and refer to \S\ref{sec:num} for numerical examples. \begin{remark} \label{Rem:UnboundedL} % Suppose that $\Psi$ satisfies $\cond{M}$, $\cond{X}$, and $\Psi\at{\pm1}=0$, but violates \cond{G} because there is some $u_\ast$ with $\abs{u_\ast}<{\Gamma}$ and $\Psi\at{u_\ast}<0$. Then $\mathcal{L}$ is unbounded from below. \end{remark} \begin{proof} We define a sequence $\at{W_n}_n\subset\mathcal{H}$ of piecewise linear profiles by $W_n\at{{\varphi}}=u_\ast$ for $\abs{{\varphi}}\leq{n}$ and $W_n\at{{\varphi}}=\mathrm{sgn}{\varphi}$ for $\abs{{\varphi}}\geq{n+1/{\Gamma}}$. By construction, $W_n-\mathcal{A}^2{W}_n$ is supported in $I_n\cup\at{-I_n}$ with $I_n=\ccinterval{n-1}{n+1+{\Gamma}}$, and a direct calculation shows \begin{align*} \mathcal{L}\at{W_n}\leq{C}+2\at{n-1}\Psi\at{u_\ast}\xrightarrow{n\to\infty}-\infty, \end{align*} where $C$ is some constant independent of $n$. \end{proof}% \subsection{Separation of phases for sequences with bounded $\mathcal{P}$}\label{sec:proof:sop}% To characterize the qualitative properties of a profile $U=\mathcal{A}{W}$ with $W\in\mathcal{C}$ we interpret $U<\tfrac{1}{2}$ and $U>\tfrac{1}{2}$ as \emph{negative phase} and \emph{positive phase}, respectively, and regard intervals in which $U$ takes intermediate values as \emph{transition layers}. Obviously, adjacent plateaus of different height are separated by a transition layer and each $U$ must exhibit at least one transition layer as it connects $-1$ to $+1$. \par% We next exploit the uniform $\fspace{L}^{\infty}$-bound for $W\in\mathcal{C}$ to derive a lower bound for the $\mathcal{P}$-contribution of each transition layer. To this end we introduce \begin{align*} Z_U=\{\bar{\varphi}\;:\;\abs{U\at{\bar{\varphi}}}\leq\tfrac{1}{2}\}. \end{align*} which is nonempty, closed and bounded as $U\in\mathcal{C}$ is continuous with $U\at{\varphi}\to\pm1$ as ${\varphi}\to\pm\infty$. \begin{remark} \label{Rem:MinCostOfZero} There exist constants $\bar{\eta}>0$ and $\bar{\mu}>0$ such that \begin{align*} \int\limits_{\bar{\varphi}-\bar\eta}^{\bar{\varphi}+\bar\eta} \Psi\at{U\at{\varphi}}\dint{\varphi} >% \bar\mu \end{align*} for each $U\in\mathcal{C}$ and $\bar{{\varphi}}\in{Z}_U$. \end{remark} \begin{proof} This follows since $U\in\mathcal{C}$ implies \begin{math} \abs{U\at{{\varphi}_2}-U\at{{\varphi}_1}}\leq% \int_{{\varphi}_2}^{{\varphi}_1}\abs{U^\prime\at{\varphi}}\dint{\varphi} \leq% 2\Gamma\abs{{\varphi}_2-{\varphi}_1} \end{math}. % In particular, we have $\abs{U\at{{\varphi}}}\leq3/4$ for all $\abs{{\varphi}-\bar{\varphi}}\leq\bar\eta=1/\at{8{\Gamma}}$ and the claim follows with $\bar\mu=\bar\sigma/\at{2\bar\eta}$ and $\bar\sigma=\sup_{\abs{u}\leq3/4}\Psi\at{u}>0$. \end{proof} In order to show that each function $U$ possesses a finite number of transition layers we introduce the following definition. A \emph{separation of phases} for a given profile $U\in\mathcal{C}$ is a finite collection of closed intervals (transition layers) $I_1,\,\tdots,\,I_m$, $m\geq1$, such that \begin{enumerate} \item the intervals are disjoint and ordered, i.e., \begin{align*} \min{I_{1}}<\max{I_{1}}<\min{I_2}<\tdots<% \max{I_{m-1}}<\min{I_{m}}<\max{I_{m}}, \end{align*} \item $Z_U$ is contained in $I_1\cup\tdots\cup{I_m}$. \item for each interval $I_j$ there exists $\bar{\varphi}_j\in{Z_U}$ such that with $\ccinterval{\bar{\varphi}_j-\bar\eta}{\bar{\varphi}_j+\bar\eta}\subset{I_j}$, % \end{enumerate} \begin{lemma} \label{Lem:ExSepPhases} Let $W\in\mathcal{C}$ be given and set $U=\mathcal{A}{W}$. Then there exists a separation of phases $I_1,\,\tdots,\,I_m$ for $U$ with \begin{align*} m\leq\mathrm{floor}\at{\mathcal{P}\at{W}/\bar\mu}\quad\text{and}\quad 2\bar\eta\leq\abs{I_j}\leq4\mathcal{P}\at{W}\bar{\eta}/\bar\mu \end{align*} for all $j=1\tdots{m}$. \end{lemma} \begin{proof} We define a finite number of points $\bar{{\varphi}}_j\in{Z_U}$ and intervals $I_j=\oointerval{\bar{{\varphi}}_j-2\bar\eta}{\bar{{\varphi}}_j-2\bar\eta}$ iteratively as follows: $\bar{{\varphi}}_1$ is the smallest element of $Z_U$, i.e. $U\at{\varphi}<-\tfrac{1}{2}$ for all ${\varphi}<\bar{{\varphi}}_1$. If $Z_U\setminus{I_1}$ is empty we stop the iteration; otherwise we choose $\bar{{\varphi}}_2$ to be the smallest zero of $Z_U$ outside of $I_1$. Then we have $\bar{{\varphi}}_2-\bar{{\varphi}}_1\geq2\bar\eta$, so Remark \ref{Rem:MinCostOfZero} yields \begin{align*} \mathcal{P}\at{W}\geq \int\limits_{\bar{{\varphi}}_1-\bar\eta}^{\bar{{\varphi}}_1+\bar\eta} \Psi\at{U\at{\varphi}}\dint{\varphi}+ \int\limits_{\bar{{\varphi}}_2-\bar\eta}^{\bar{{\varphi}}_2+\bar\eta} \Psi\at{U\at{\varphi}}\dint{\varphi} \geq{2}\bar\mu. \end{align*} If possible, we now define $\bar{{\varphi}}_3$ as the minimum of $Z_U\setminus\at{I_1\cup{I_2}}$ and proceed iteratively until the iteration stops after $m\leq\mathrm{floor}\at{\mathcal{P}\at{W}/\bar\mu}$ steps. By construction, we have $Z_U\subset\bigcup_{j=1}^m{I_j}$ and $\abs{I_j}\leq4\bar\eta$ for all $j$ but the intervals $I_j$ may overlap. Finally, we obtain the desired separation of phases by merging overlapping intervals. \end{proof} Our main result in this section concerns sequences $\at{W_n}_n\in\mathcal{C}$ with bounded $\mathcal{P}\at{W_n}$. It guarantees, roughly speaking, the existence of \emph{compatible} transitions layers which $\at{i}$ have the same length, $\at{ii}$ separate the same phases, and $\at{iii}$ depart from each other. For an illustration we refer to Figure \ref{fig:sep_phase}. \begin{lemma} \label{Lem:SepPhasesForSequ} Let $\at{W_n}_{n}\subset\mathcal{C}$ be a sequence with $\limsup_{n\to\infty}\mathcal{P}\at{W_n}<\infty$. Then there exists a not relabelled subsequence along with a finite number of intervals $J_1,\,\tdots,\,J_m$ with the following properties: \begin{enumerate} \item Each interval $J_j$ is centred around zero, i.e., $J_j=\ccinterval{-\eta_j}{\eta_j}$ for some $0<\eta_j<\infty$. \item For each $n$ there exist shifts ${\varphi}_{n,1}<\tdots<{\varphi}_{n,m}$ such that \begin{enumerate} \item ${\varphi}_{n,1}+J_1,\,\tdots,\,{\varphi}_{n,m}+J_m$ is a separation of phases for $U_n=\mathcal{A}{W_n}$, \item ${\varphi}_{n,j+1}-{\varphi}_{n,j}\to\infty$ as $n\to\infty$ for all $j=1{\tdots}m-1$. \end{enumerate} \item There exists a choice of signs $s_0,\,\tdots,\,s_{m}$ with $s_0=-1$, $s_m=1$, and $s_j\in\{-1,\,+1\}$ such that \begin{align*} \begin{array}{lclclclcl} \mathrm{sgn}\at{U_n\at{\varphi}}&=&s_0 &\quad\text{for}\quad&% &&{\varphi}&<&{\varphi}_{n,1}-\eta_{1} \\% \mathrm{sgn}\at{U_n\at{\varphi}}&=&s_1 &\quad\text{for}\quad&% {\varphi}_{n,1}+\eta_1&<&{\varphi}&<&{\varphi}_{n,2}-\eta_{2} \\% &&&\tdots&&&&& \\% \mathrm{sgn}\at{U_n\at{\varphi}}&=&s_{m-1} &\quad\text{for}\quad&% {\varphi}_{n,m-1}+\eta_{m-1}&<&{\varphi}&<&{\varphi}_{n,m}-\eta_m \\% \mathrm{sgn}\at{U_n\at{\varphi}}&=&s_{m+1} &\quad\text{for}\quad&% {\varphi}_{n,m}+\eta_m&<&{\varphi}&& \end{array} \end{align*} hold for all $n$. \end{enumerate} \end{lemma} \begin{proof} Thanks to Lemma \ref{Lem:ExSepPhases} we can extract a subsequence such that each $U_n$ has a separation of phases that consists of $m$ intervals $I_{n,1},\tdots,{I_{n,m}}$ where $m$ is independent of $n$. We denote the centre of $I_{n,j}$ by ${\varphi}_{n,j}$, set $J_{n,j}=I_{n,j}-{\varphi}_{n,j}$, and notice that $\ul{c}\leq\abs{J_{n,j}}\leq\ol{c}$ and ${\varphi}_{n,j+1}-{\varphi}_{n,j}>\ul{c}$ for some constants $\ol{c}>\ul{c}>0$ independent of $n$ and $j$. In particular, the intervals $J_{j}=\bigcup_{n}J_{n,j}$ have finite length $\ol{c}\leq\abs{J_j}\leq\ol{c}$. \par Our strategy for the proof is to refine $m$, the subsequence $\at{W_n}_n$, the phase shifts ${\varphi}_{n,j}$, and the intervals $J_j$ in several steps. To this end we start the following algorithm at level $1$. \begin{enumerate} \item[]% Level k \begin{enumerate} \item[] If $m=k$, then we stop the algorithm. \item[] If ${\varphi}_{n,k+1}-{\varphi}_{n,k}\to\infty$ as $n\to\infty$ along a subsequence, then we extract this subsequence and jump to level $k+1$. \item[] % If $m<k$ and $0<\sup_{n}\at{{\varphi}_{n,k+1}-{\varphi}_{n,k}}<\infty$, then we merge $J_k$ and $J_{k+1}$ as follows: At first we choose $\tilde{J}_{k}$ sufficiently large such that \begin{align*} \tilde{J}_{k}\supseteq{J}_{k}\cup% {\bigcup_{n}\at{{\varphi}_{n,k+1}-{\varphi}_{n,k}+{J}_{k+1}}}% \end{align*} Secondly we define $\tilde{{\varphi}}_{n,k}={\varphi}_{n,k}$, $\tilde{m}=m-1$ and \begin{align*} \begin{array}{lclclclcll} \tilde{J}_j&=&J_j&\quad\text{and}\quad&\tilde{{\varphi}}_{n,j} &=&{\varphi}_{n,j}&\quad\text{for}\quad&{j<k},\\ \tilde{J}_j&=&J_{j+1}&\quad\text{and}\quad&\tilde{{\varphi}}_{n,j}&=& {\varphi}_{n,j+1}&\quad\text{for}\quad&{j>k}. \end{array} \end{align*} Finally we restart level $k$ with $\tilde{m}$, $\tilde{{\varphi}}_{n,j}$, and $\tilde{J}_j$ instead of $m$, ${\varphi}_{n,j}$, and $J_j$. \end{enumerate} \end{enumerate} This algorithm stops after a finite number of steps when $m-k=0$. It provides intervals $J_j$ and phase shifts ${\varphi}_{n,j}$ for $j=1\tdots{m}$ and $n\in{N}$ with $\lim_{n\to\infty}{\varphi}_{n,j+1}-{\varphi}_{n,j}=\infty$ for all $j$. By extracting subsequences we can also ensure that, for each $n$, the intervals ${\varphi}_{n_j}+J_{n,j}$ are pairwise disjoint and provide therefore a separation of phases for $U_n$. Finally, by extracting further subsequences if necessary we guarantee the existence of a choice of signs. \end{proof} \subsection{Existence of minimizers for $\mathcal{L}$}% We now finish the existence proof for fronts. \begin{theorem} \label{Theo:Minimiser}% $\mathcal{L}$ attains its minimum on $\mathcal{C}$ and each minimizer is a front. \end{theorem} \begin{proof} \emph{\underline{Step 0.}} % We start with some notations. For a given minimizing sequence $\at{W_n}_n\subset\mathcal{C}$ we define \begin{align*} U_n=\mathcal{A}{W_n},\qquad{S_n}=\mathrm{sgn}{U_n}, \end{align*} and for each $K>0$ we introduce the operator \begin{align*} E_K:\mathcal{C}\to\mathcal{C},\quad E_KW=\chi_{\ccinterval{-K}{K}}\at{W-W_{\rm sh}}+W_{\rm sh}. \end{align*} Here $\chi_{\ccinterval{-K}{K}}$ denotes the usual indicator function, so we have $\norm{E_KW-W}_2\to0$ as $K\to\infty$ for each $W\in\mathcal{C}$. Finally, within this proof $C$ always denotes a positive constant that is independent of $n$ and $K$, but the value of $C$ may change from line to line. \par \emph{\underline{Step 1.}} By assumption and $\mathcal{M}\at{W_n-W_{\rm sh}}\geq0$ we have \begin{align} \label{Theo:Minimiser.ZEqn1} \mathcal{P}\at{W_n} =% {\mathcal{L}\at{W_n}}-{\mathcal{N}\at{W_n}} \leq% {C}+\norm{W-W_{\rm sh}}_{\infty}\norm{W_{\rm sh}-\mathcal{A}^2W_{\rm sh}}_1 \leq{C}. \end{align} Therefore we can extract (a not relabelled) subsequence for which Lemma \ref{Lem:SepPhasesForSequ} provides a finite number of intervals $J_1\tdots{J}_m$, sequences of phase shifts $\at{{\varphi}_{n,\,j}}_n$ and a choice of signs $\at{s_{n,\,j}}_n$. There exits at least one $1\leq{j_\ast}\leq{m}$ such that $s_{j_\ast-1}=-1$ and $s_{j_\ast}=+1$, and since $\mathcal{L}$ is invariant under shifts we can assume that ${\varphi}_{n,\,j_\ast}=0$. With $J_{j_\ast}=\ccinterval{-\eta_{j_\ast}}{\eta_{j_\ast}}$ and due to $\lim_{n\to\infty}{\varphi}_{n,\,j+1}-{\varphi}_{n,\,j}=\infty$ we then have \begin{align*} \limsup\limits_{n\to\infty} U_{n}\at{\varphi}\leq-\tfrac{1}{2} \quad\text{for}\quad {\varphi}<-\eta_{j_\ast}, \qquad \liminf\limits_{n\to\infty}U_{n}\at{\varphi}\geq\tfrac{1}{2} \quad\text{for}\quad {\varphi}>\eta_{j_\ast}. \end{align*} By compactness we can extract a further subsequence such that $W_n\rightharpoonup{W_\infty}$ weakly$\star$ in $\fspace{W}^{1,\,\infty}$. In particular, $W_n$ converges to $W_\infty$ uniformly on each compact interval, and hence \begin{align} \label{Theo:Minimiser.EqnA1}% E_K{W_n}\xrightarrow{n\to\infty}E_KW_\infty ,\quad% E_K{U_n}\xrightarrow{n\to\infty}E_KU_\infty \quad% \text{strongly in $\fspace{L}^2$ for all $K$}. \end{align} \par \emph{\underline{Step 2.}} % Towards $W_\infty\in\mathcal{C}$ we show that $W_n-S_n$ is uniformly bounded in $\fspace{L}^2$. The first observation is that \eqref{Theo:Minimiser.ZEqn1} combined with \eqref{Eqn:LowerEstimateForG} implies \begin{align} \label{Theo:Minimiser.EqnB1} \norm{U_n-S_n}_2\leq{C}. \end{align} The second observation is that both $S_n-{\mathcal{A}}S_n$ and $S_n-\mathcal{A}^2S_n$ are supported in the $1$-neighbourhood of $I_n=\bigcup_{j=1}^m\at{J_j+{\varphi}_{n,\,j}}$. Therefore, $\abs{I_n}\leq\sum_{j=1}^{m}\abs{J_j}\leq{C}$ yields \begin{align*} \norm{S_n-{\mathcal{A}}S_n}_2\leq{C} ,\qquad \norm{S_n-\mathcal{A}^2S_n}_1\leq{C} ,\qquad% \abs{\mathcal{N}\at{S_n}}=\abs{\skp{S_n-\mathcal{A}^2S_n}{S_n}}\leq{C}, \end{align*} and by \eqref{Theo:Minimiser.EqnB1} we find \begin{align} \label{Theo:Minimiser.EqnB2}% \norm{U_n-{\mathcal{A}}S_n}_2\leq\norm{U_n-S_n}_2+\norm{S_n-{\mathcal{A}}S_n}_2\leq{C}. \end{align} Exploiting \eqref{Lem:NProps.Eqn6} for $W_2=W_n$ and $W_1=S_n$ gives \begin{align*} \mathcal{M}\at{W_n-S_n} &=% {\mathcal{N}\at{W_n}}-\mathcal{N}\at{S_n}-\skp{W_n-S_n}{S_n-\mathcal{A}^2S_n} \\&\leq% \mathcal{L}\at{W_n}+\abs{\mathcal{N}\at{S_n}}+ \at{\norm{W_n}_\infty+\norm{S_n}_\infty}\norm{S_n-\mathcal{A}^2S_n}_1\leq{C}, \end{align*} and with \eqref{Theo:Minimiser.EqnB2} we obtain \begin{align} \label{Theo:Minimiser.EqnB3}% \norm{W_n-S_n}^2_2&=\mathcal{M}\at{W_n-S_n}+\norm{U_n-{\mathcal{A}}S_n}^2_2 \leq{C}. \end{align} Now we are able to show $W_\infty\in\mathcal{C}$. From \eqref{Theo:Minimiser.EqnB3} we infer that \begin{align*} \norm{E_KW_n-W_{\rm sh}}_2 \leq\norm{E_KW_n-E_KS_n}_2+\norm{E_K{S_n}-E_KW_{\rm sh}}_2\leq{C}+\norm{E_K{S_n}-E_KW_{\rm sh}}_2, \end{align*} and with $S_n\at{{\varphi}}\to{W_{\rm sh}}\at{\varphi}$ as $n\to\infty$ for all ${\varphi}\neq{J}_{j_\ast}$ we find \begin{align*} \norm{E_KW_\infty-W_{\rm sh}}_2\leq{C}. \end{align*} Passing to the limit $K\to\infty$ now gives $W\in\mathcal{H}$, and $W\in\mathcal{C}$ follows because $W_\infty$ was defined as weak$\star$ limit in $\fspace{W}^{1,\,\infty}$. \par \emph{\underline{Step 3.}} % There remains to show that $W_\infty$ minimizes $\mathcal{L}$. From \eqref{Lem:NProps.Eqn6} we infer that \begin{align} \label{Theo:Minimiser.EqnC1} \mathcal{N}\at{W_n} =% \mathcal{N}\at{E_KW_n}+\mathcal{M}\at{W_n-E_KW_n}+ \skp{W_n-E_KW_n}{E_KW_n-\mathcal{A}^2\at{E_KW_n}} \end{align} and \begin{align} \label{Theo:Minimiser.EqnCa} \begin{split} \abs{\mathcal{N}\at{E_K{W}_n}-\mathcal{N}\at{E_K{W}_\infty}} \leq&\quad% \mathcal{M}\at{E_KW_n-E_KW_\infty}\\&+ \norm{E_KW_n-E_KW_\infty}_2\norm{E_KW_\infty-\mathcal{A}^2\at{E_KW_\infty}}_2 \end{split} \end{align} hold for all $K$ and $n\in{\mathbb{N}}\cup\{\infty\}$. Combining \eqref{Theo:Minimiser.EqnCa} with \eqref{Theo:Minimiser.EqnA1} gives \begin{align*} \mathcal{N}\at{E_K{W}_n}\xrightarrow{n\to\infty}\mathcal{N}\at{E_K{W}_\infty}. \end{align*} Moreover, since $E_K{W_n}-\mathcal{A}^2\at{E_K{W_n}}$ is supported in $\ccinterval{-K-1}{K+1}$ for all $n$, we also have \begin{align*} \skp{W_n-E_KW_n}{E_KW_n-\mathcal{A}^2\at{E_KW_n}} \xrightarrow{n\to\infty}% \skp{W_\infty-E_KW_\infty}{E_KW_\infty-\mathcal{A}^2\at{E_KW_\infty}}, \end{align*} and passing to the limit $n\to\infty$ in \eqref{Theo:Minimiser.EqnC1} provides \begin{align} \label{Theo:Minimiser.EqnC2} \liminf\limits_{n\to\infty}\mathcal{N}\at{W_n} \geq% \mathcal{N}\at{E_KW_\infty}+ \skp{W_\infty-E_KW_\infty}{E_KW_\infty-\mathcal{A}^2\at{E_KW_\infty}}, \end{align} where we used that $\mathcal{M}\at{W_n-E_KW_n}\geq0$ according to Lemma \ref{Lem:MProps}. On the other hand, evaluating \eqref{Theo:Minimiser.EqnC1} for $n=\infty$ gives \begin{align*} \mathcal{N}\at{W_\infty}=\mathcal{N}\at{E_KW_\infty}+\mathcal{M}\at{W_\infty-E_KW_\infty}+ \skp{W_\infty-E_KW_\infty}{E_KW_\infty-\mathcal{A}^2\at{E_KW_\infty}} \end{align*} and due to $\mathcal{M}\at{W_\infty-E_KW_\infty}\to0$ as $K\to\infty$ we find \begin{align} \label{Theo:Minimiser.EqnC3} \mathcal{N}\at{E_KW_\infty}+ \skp{W_\infty-E_KW_\infty}{E_KW_\infty-\mathcal{A}^2\at{E_KW_\infty}} \xrightarrow{K\to\infty}\mathcal{N}\at{W_\infty}. \end{align} The combination of \eqref{Theo:Minimiser.EqnC2} and \eqref{Theo:Minimiser.EqnC3} reveals \begin{align} \label{Theo:Minimiser.EqnC4} \liminf\limits_{n\to\infty}\mathcal{N}\at{W_n}\geq\mathcal{N}\at{W_\infty}, \end{align} and Fatou's Lemma provides \begin{align} \label{Theo:Minimiser.EqnC5} \liminf\limits_{n\to\infty}\mathcal{P}\at{W_n}\geq\mathcal{P}\at{W_\infty} \end{align} due to $\Psi\geq0$ and since $W_n$ converges to $W_\infty$ pointwise. Adding \eqref{Theo:Minimiser.EqnC4} and \eqref{Theo:Minimiser.EqnC5} we conclude that $W_\infty$ is in fact a minimizer of $\mathcal{L}$, and Lemma \ref{Lem:FrontsAndMinimisers} guarantees that $W_\infty$ solves the front equation \eqref{Norm.Problem.Eqn}. \end{proof} We conclude with some remarks. \begin{enumerate} \item Theorem \ref{Intro:MainTheo} follows by combining Lemma \ref{Lem:JumpCond}, Lemma \ref{Lem:Normalisation}, Lemma \ref{Lem:FrontsAndMinimisers}, Remark \ref{Rem:UnboundedL}, and Theorem \ref{Theo:Minimiser}. \item The assertions of Theorem \ref{Theo:Minimiser} can be sharpened as follows. For each minimizing sequence $\at{W_n}_n$ we have equality signs in both \eqref{Theo:Minimiser.EqnC4} and \eqref{Theo:Minimiser.EqnC5}, and hence \begin{align*} \lim\limits_{K\to\infty}\lim\limits_{n\to\infty}\mathcal{M}\at{W_n-E_KW_n}=0. \end{align*} This implies that there is only one interval $J_1=J_{j_\ast}$ and in turn that $W_n-W_\infty\to0$ strongly in $\fspace{L}^2$. In this sense each minimizing sequence obeys exactly one transition from negative phase to positive phase. \item If $\Phi^\prime$ is increasing in $\ccinterval{-1}{+1}$ we can improve the existence result for fronts as follows. We choose ${\Gamma}=1$ in \cond{I} and consider the set \begin{align*} \wt{\mathcal{C}}=\{W\in\mathcal{C}\;:\;W^\prime\geq0\}. \end{align*} Then $\wt{\mathcal{C}}$ is an invariant set for the gradient flow of $\mathcal{L}$ and again one can show that $\mathcal{L}$ restricted to $\wt{\mathcal{C}}$ attains it minimum (a proof tailored to monotone profiles is given in \cite{HR10b}). In particular, in this case there exist action minimizing fronts with monotone profile $W$. \end{enumerate} \section{Approximation of fronts}\label{sec:num} In this section we illustrate the analytical results from \S\ref{sec:proof} by some numerical simulations. To this end we discretize the Euler scheme for the gradient flow of $\mathcal{L}$, see \eqref{Eqn:EulerScheme}, as follows: \begin{samepage} \begin{enumerate} \item Fix a finite interval $\ccinterval{-L}{+L}$ and introduce equidistant grid points by ${\varphi}_k=-L+2kL/D$, where $k=0\tdots{D}$ and $D\in{\mathbb{N}}$ is large. \item Approximate each profile $W\in\mathcal{C}$ by the discrete vector $W_i=W\at{{\varphi}_i}$ and impose the boundary conditions $W_i=-1$ and $W_i=+1$ for $i<0$ and $i>D$, respectively. \item Replace the integrals in the definition of $\mathcal{A}$ by Riemann sums with respect to the ${\varphi}_i$'s. \item Choose ${\lambda}$ sufficiently small and initialize the iteration \eqref{Eqn:EulerScheme} with shock initial data $W_i=\mathrm{sgn}{{\varphi}_i}$. \end{enumerate} \end{samepage} \begin{figure}[ht!]% \centering{% \begin{minipage}[c]{0.285\textwidth}% \includegraphics[width=\textwidth, draft=\figdraft]% {\figfile{ex_1_force}}% \end{minipage}% \hspace{0.025\textwidth}% \begin{minipage}[c]{0.285\textwidth}% \includegraphics[width=\textwidth, draft=\figdraft]% {\figfile{ex_1_prof}}% \end{minipage}% \hspace{0.025\textwidth}% \begin{minipage}[c]{0.315\textwidth}% \includegraphics[width=\textwidth, draft=\figdraft]% {\figfile{ex_1_res}}% \end{minipage}% \\% \begin{minipage}[c]{0.285\textwidth}% \includegraphics[width=\textwidth, draft=\figdraft]% {\figfile{ex_2_force}}% \end{minipage}% \hspace{0.025\textwidth}% \begin{minipage}[c]{0.285\textwidth}% \includegraphics[width=\textwidth, draft=\figdraft]% {\figfile{ex_2_prof}}% \end{minipage}% \hspace{0.025\textwidth}% \begin{minipage}[c]{0.315\textwidth}% \includegraphics[width=\textwidth, draft=\figdraft]% {\figfile{ex_2_res}}% \end{minipage}% \\% \begin{minipage}[c]{0.285\textwidth}% \includegraphics[width=\textwidth, draft=\figdraft]% {\figfile{ex_3_force}}% \end{minipage}% \hspace{0.025\textwidth}% \begin{minipage}[c]{0.285\textwidth}% \includegraphics[width=\textwidth, draft=\figdraft]% {\figfile{ex_3_prof}}% \end{minipage}% \hspace{0.025\textwidth}% \begin{minipage}[c]{0.315\textwidth}% \includegraphics[width=\textwidth, draft=\figdraft]% {\figfile{ex_3_res}}% \end{minipage}% }% \caption{% Three examples with admissible potential as in Assumption \ref{Ass:Pot} where $\Phi^\prime$ is plotted in the invariant interval. During the iteration the profiles $W$ converge to a front. }% \label{Fig:Num1}% \end{figure}% \begin{figure}[ht!]% \centering{% \begin{minipage}[c]{0.285\textwidth}% \includegraphics[width=\textwidth, draft=\figdraft]% {\figfile{ex_1_pot}}% \end{minipage}% \hspace{0.025\textwidth}% \begin{minipage}[c]{0.285\textwidth}% \includegraphics[width=\textwidth, draft=\figdraft]% {\figfile{ex_2_pot}}% \end{minipage}% \hspace{0.025\textwidth}% \begin{minipage}[c]{0.285\textwidth}% \includegraphics[width=\textwidth, draft=\figdraft]% {\figfile{ex_3_pot}}% \end{minipage}% \\% \begin{minipage}[c]{0.285\textwidth}% \includegraphics[width=\textwidth, draft=\figdraft]% {\figfile{ex_4_pot}}% \end{minipage}% \hspace{0.025\textwidth}% \begin{minipage}[c]{0.285\textwidth}% \includegraphics[width=\textwidth, draft=\figdraft]% {\figfile{ex_5_pot}}% \end{minipage}% \hspace{0.025\textwidth}% \begin{minipage}[c]{0.285\textwidth}% \includegraphics[width=\textwidth, draft=\figdraft]% {\figfile{ex_6_pot}}% \end{minipage}% }% \caption{% Potentials $\Phi$ with front parabolas for the examples from Figure \ref{Fig:Num1} (top row) and \ref{Fig:Num3} (bottom row). }% \label{Fig:Num2}% \end{figure}% \begin{figure}[ht!]% \centering{% \begin{minipage}[c]{0.285\textwidth}% \includegraphics[width=\textwidth, draft=\figdraft]% {\figfile{ex_4_force}}% \end{minipage}% \hspace{0.025\textwidth}% \begin{minipage}[c]{0.3\textwidth}% \includegraphics[width=\textwidth, draft=\figdraft]% {\figfile{ex_4_prof_1}}% \end{minipage}% \hspace{0.025\textwidth}% \begin{minipage}[c]{0.3\textwidth}% \includegraphics[width=\textwidth, draft=\figdraft]% {\figfile{ex_4_prof_2}}% \end{minipage}% \\% \begin{minipage}[c]{0.285\textwidth}% \includegraphics[width=\textwidth, draft=\figdraft]% {\figfile{ex_5_force}}% \end{minipage}% \hspace{0.025\textwidth}% \begin{minipage}[c]{0.3\textwidth}% \includegraphics[width=\textwidth, draft=\figdraft]% {\figfile{ex_5_prof_1}}% \end{minipage}% \hspace{0.025\textwidth}% \begin{minipage}[c]{0.3\textwidth}% \includegraphics[width=\textwidth, draft=\figdraft]% {\figfile{ex_5_prof_2}}% \end{minipage}% \\% \begin{minipage}[c]{0.285\textwidth}% \includegraphics[width=\textwidth, draft=\figdraft]% {\figfile{ex_6_force}}% \end{minipage}% \hspace{0.025\textwidth}% \begin{minipage}[c]{0.3\textwidth}% \includegraphics[width=\textwidth, draft=\figdraft]% {\figfile{ex_6_prof_1}}% \end{minipage}% \hspace{0.025\textwidth}% \begin{minipage}[c]{0.3\textwidth}% \includegraphics[width=\textwidth, draft=\figdraft]% {\figfile{ex_6_prof_2}}% \end{minipage}% }% \caption{% Three counterexamples for the existence of action minimizing fronts. The first two examples satisfy the macroscopic constraints but violate the graph condition \cond{G}, so the iteration minimizes the action via an extending plateau. The third example violates the constraint $\Phi\at{+1}=\Phi\at{-1}$, so the profiles converge to one of the asymptotic states. }% \label{Fig:Num3}% \end{figure}% In numerical simulations, the resulting iteration scheme has good convergence properties and decreases the action provided that ${\lambda}$ is sufficiently small and $L$ is sufficiently large. Three examples with normalized front data and $\Phi$ as in Assumption \ref{Ass:Pot} are shown in Figure \ref{Fig:Num1}, where $\Phi^\prime$ is always plotted over the invariant interval $\ccinterval{-{\Gamma}}{{\Gamma}}$. For plots of $\Phi$ see Figure \ref{Fig:Num2}. In the first example $\Phi^\prime$ is increasing in $\ccinterval{-1}{+1}$ and the front profile $W$ turns out to be monotone. We refer to the remark at the end of \S\ref{sec:proof} for an explanation, and to \cite{HR10b} for more examples. The other two examples in Figure \ref{Fig:Num1} illustrate that the front profiles for non-convex $\Phi$ are in general non-monotone. \par\quad\newline\noindent In order to illustrate the necessity of the graph condition \cond{G} we present in Figure \ref{Fig:Num3} simulations for potentials that violate this condition. In the first example the interval $\ccinterval{-1}{1}$ is invariant under $\Phi^\prime$ but $\Psi$ is negative in this interval. After some initial iterations the profiles exhibit an extending plateau at $-1<w_\ast<1$, where $w_\ast$ is the minimizer of $\Psi$ in $\ccinterval{-1}{1}$ and satisfies $w_\ast=\Phi^\prime\at{w_\ast}$. The onset of the extending plateau is a direct consequence of the energy landscape of $\mathcal{L}$, see Remark \ref{Rem:UnboundedL}. The second simulation provides another counterexample for the existence of action minimizing fronts. Here $\Psi$ has the correct sign close to $\pm1$ but attains again a negative minimum in $-1<w_\ast<1$. As before, the profiles minimize their action by converging to the `global minimizer' $W\equiv{w_\ast}$. Finally, the third example illustrates what happens if the asymptotic states violate the macroscopic constraints. More precisely, here we violate $\eqref{Lem:JumpCond.Eqn2}$ due to $\Phi\at{+1}>\Phi\at{-1}$, and observe that the profiles converge to the global minimizer $W\equiv+1$. \section*{Acknowledgements}% \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace} \providecommand{\MR}{\relax\ifhmode\unskip\space\fi MR } \providecommand{\MRhref}[2]{% \href{http://www.ams.org/mathscinet-getitem?mr=#1}{#2} } \providecommand{\href}[2]{#2}
\section{Introduction} Recent discovery of the two-dimensional (2D) quantum spin Hall system\cite{Mele,Bernevig,wu2006,xu2006,Fu,Qi,Bernevig2,Konig}, and its three-dimensional (3D) generalization, dubbed topological insulator, \cite{Moore,Fu2,Teo,Hsieh,Qi} has established a new state of matter in the time-reversal symmetric systems. The topological order in the bulk with the gap guarantees the presence of the one-dimensional (1D) channels along the edge of the 2D sample, or the 2D metal on the surface of the 3D sample. These edge and surface states are protected by the time-reversal symmetry and the topology of the bulk gap, and are robust against the disorder scattering and electron-electron interactions. In topological surface state on 3D topological insulator, the electrons obey the 2D Dirac equations. This has been beautifully demonstrated by the spin- and angle-resolved photoemission spectroscopy~\cite{Hasan,Ando}. To date, various interesting properties of the topological insulator have been predicted, particularly those relevant to magnetism~\cite{Maciejko,Qi2,Liu,Yokoyama,Tanaka} and electron-electron interaction\cite{Hou,Strom,Tanaka2,Teo2,Raghu,Zhang}. Since the surface state of topological insulator has a suppressed backward scattering for nonmagnetic impurities, this is a promising material for designing the quantum curcuit. Therefore, electric control of transport on the surface of topological insulator is an important issue. In fact, the chiral edge channels are predicted to appear at the interface of two ferromagnets with magnetization along $z$ and $-z$ directions, i.e., perpendicular to the surface.\cite{Niemi} Also, scanning tunneling microscopy experiments and related theories have revealed the electronic states due to the impurity scattering or the terrace edge on the surface of topological insulator\cite{Alpichshev,XZhang,Lee,Wang,Biswas,Biswas2}. In this paper, we study the 1D states that are formed on the topological surface under the gate electrode. The energy dispersion of these states is almost linear in the momentum with the velocity sensitively depending on the strength of the gate voltage. The energy is also restricted to be positive or negative depending on the strength of the gate voltage. Consequently, the local density of states near the gated region has an asymmetric structure with respect to zero energy. With the electron-electron interaction, the correlation effect can be tuned by the gate voltage. We also suggest a tunneling experiment to verify the presence of these bound states. \begin{figure}[htb] \begin{center} \scalebox{0.8}{ \includegraphics[width=8.0cm,clip]{fig1.eps} } \end{center} \caption{(Color online) (Upper) schematics of the proposed setup. The gate electrode is attached on the surface of the topological insulator where 1D helical mode is generated. A magnetic impurity placed on the gated region produces the backward scattering. (Lower) dispersions of surface and bound states. } \label{fig1} \end{figure} \begin{figure}[htb] \begin{center} \scalebox{0.8}{ \includegraphics[width=8.0cm,clip]{fig3.eps}} \end{center} \caption{(Color online) The dispersion of the bound states as a function of $k_y$ for various $U/E_F$ with $k_F L =0.1$ and $E_F = v_F k_F$. } \label{fig3} \end{figure} We consider 2D topological surface on which the 1D gate electrode is attached as shown in Fig. \ref{fig1}. The gate electrode is attached in the region $0 \le x \le L$. The interface between gated and ungated regions is parallel to the $y$-axis. The electrons on the topological surface obey the 2D Dirac Hamiltonian $H = v_F \bm{k} \cdot \bm{\sigma}$ and under the gate, we add the potential $U$, namely $H = v_F \bm{k}' \cdot \bm{\sigma} + U$.\cite{note} Here, $v_F$ is the Fermi velocity, $\bm{k} (\bm{k'})$ is a wavevector and $\bm{\sigma}$ is a vector of Pauli matrices. The dispersion relation is $E = \pm v_F \sqrt {k_x^2 + k_y^2 } = \pm v_F \sqrt {k_x^{'2} + k_y^2 } + U.$ Note that due to the translational symmetry along the $y$-axis, $k_y$ is conserved. Now, let us regard $v_F \left| {k_y^{} } \right|$ as a gap for a fixed $k_y$ and consider the bound states formed inside this gap. By setting $k_x = i\kappa (\kappa>0)$ outside the gated region to search for the bound states near the gate, with the above Hamiltonian, we have wave functions in each region represented as \begin{eqnarray} \psi (x \le 0) = ae^{\kappa x} \left( {\begin{array}{*{20}c} E \\ {iv_F (k_y^{} - \kappa )} \\ \end{array}} \right), \\ \psi (0 \le x \le L) = be^{ik_x 'x} \left( {\begin{array}{*{20}c} {E - U} \\ {v_F (k_x ' + ik_y^{} )} \\ \end{array}} \right) \nonumber \\ + ce^{ - ik_x 'x} \left( {\begin{array}{*{20}c} {E - U} \\ {v_F ( - k_x ' + ik_y^{} )} \\ \end{array}} \right), \\ \psi (x \ge L) = de^{ - \kappa x} \left( {\begin{array}{*{20}c} E \\ {iv_F (k_y^{} + \kappa )} \\ \end{array}} \right). \end{eqnarray} In the above expressions, the common factor $e^{ik_y y} $ is omitted for simplicity. By matching the wavefunctions at the interfaces $x=0$ and $L$, we obtain \begin{eqnarray} v_F^2 k_y^2 + v_F^2 \kappa \frac{{k_x '}}{{\tan k_x 'L}} = E(E - U) \end{eqnarray} where $v_F k_x ' = \sqrt {(E - U)^2 - v_F^2 k_y^2 }$. By solving the above equation numerically, we plot the dispersion of the bound states as a function of $k_y$ for various $U/E_F$ with $k_F L =0.1$ and $E_F = v_F k_F$ in Fig. \ref{fig3}. As seen, the bound states are generated with an almost linear dispersion in the momentum. The sign and the velocity of this dispersion strongly depend on the potential parameter $U$. In order to make the following discussion more transparent, let us take the limit $U \to \infty$ and $L \to 0$ while keeping $Z \equiv UL = const$. Then, we have $v_F \kappa = - E\tan Z (> 0)$, and hence $E = \pm v_F (\cos Z)k_y$ and $\kappa = \mp (\sin Z)k_y$. When the sign of $\tan Z$ is fixed, then that of $E$ is determined. There are two branches of the bound states, which constitute a 1D helical mode. This is a physical realization of the Tomonaga model which is characterized by two dispersions defined for positive energy only (and no dispersion for negative energy).\cite{Tomonaga} How does this helical mode manifest itself in observable quantities? In the following, we show its salient feature in local density of states and the current along the 1D channel. First, we study the local density of states near the gated region ($x \le 0$). The wavefunction of the bound states is given by \begin{eqnarray} \psi _B (x \le 0) = \sqrt {\frac{\kappa }{{1 \pm \sin Z}}} e^{\kappa x} \left( {\begin{array}{*{20}c} {\cos Z} \\ {i( \pm 1 + \sin Z)} \\ \end{array}} \right). \end{eqnarray} Note that the spins of the helical mode are opposite to each other. Namely, the expectation value of the Pauli matrices is given by $\left\langle \bm{\sigma} \right\rangle \parallel \left( {0, \pm \cos Z, \mp \sin Z} \right)^t$. The total local density of states can be calculated as \begin{eqnarray} \rho (E ,x) = \frac{1}{{4\pi ^2 }}\int {dk_x dk_y \left| {\psi _S^{} } \right|^2 \delta (E - E_S )} \nonumber \\ + \frac{1}{{2\pi }}\int {dk_y \left| {\psi _B^{} } \right|^2 } \delta (E - E_B )\\ = \rho ^{(0)} + \delta \rho (E ,x) + \rho _B (E ,x) \end{eqnarray} with $\rho ^{(0)} = \frac{{\left| E \right|}}{{2 \pi v_F^2 }}$ and \begin{widetext} \begin{eqnarray} \delta \rho (E ,x) = - \frac{E }{{ \pi ^2 v_F^2 }}\int_0^{\pi /2} {d\theta \sin ^2 \theta {\mathop{\rm Im}\nolimits} \left( {e^{ - 2i(E \cos \theta ) x/v_F} \frac{{\sin Z}}{{\cos \theta \cos Z - i{\mathop{\rm sgn}} (E )\sin Z}}} \right)} \end{eqnarray} \end{widetext} where $\psi _S$ is the wavefunction for the scattering state, and $E_S$ and $E_B$ are the energy of the scattering state and the bound states, respectively. $\rho ^{(0)}$ and $\delta \rho$ come from the scattering states. \cite{Biswas2} The local density of states stemming from the bound states $\rho _B$ is given by\begin{eqnarray} \rho _B (E ,x \le 0) = - \frac{E }{ \pi v_F^2 }\frac{{\tan Z}}{{\left| {\cos Z} \right|}}e^{ - 2(E \tan Z) x/v_F} \label{dosb} \end{eqnarray} for $E \tan Z <0$ and $\rho _B =0$ for $E \tan Z >0$. The $\rho ^{(0)}$ and $\delta \rho (E ,x)$ are obtained in Ref. \onlinecite{Biswas2}, while $\rho _B (E ,x)$ has been missed in the literature. \begin{figure}[htb] \begin{center} \scalebox{0.8}{ \includegraphics[width=11.0cm,clip]{fig4.eps}} \end{center} \caption{ (Color online) (a) $\rho (E, 0) E_F$, (b) $\rho (E, -1/k_F) E_F$, (c) $\rho _B (E ,0) E_F$ and (d) $\rho _B (E, -1/k_F) E_F$ for several $Z$. Note $\rho _B (E ,x)=0$ at $Z=0$ and also that the scales of (a), (c) and (b), (d) are different, and the contribution from $\rho_B$ is larger in (a) (at $x=0$) compared with (b) (at $x=-1/k_F$). } \label{fig4} \end{figure} In Fig. \ref{fig4}, we show (a) $\rho (E, 0) E_F$, (b) $\rho (E, -1/k_F) E_F$, (c) $\rho _B (E ,0) E_F$ and (d) $\rho _B (E, -1/k_F) E_F$ for several $Z$. As shown in Fig. \ref{fig4} (a) and Fig. \ref{fig4} (b), the contribution from the bound states gives an asymmetry in the local density of states: $\rho_A(E,x) \equiv \rho(E,x)-\rho(-E,x) \ne 0$ (note $\rho _B (E ,x)=0$ at $Z=0$ and also that $\rho ^{(0)} $ and $\delta \rho$ are even function of $E$). Since the bound states decay exponentially in space as seen in Eq.(\ref{dosb}), the contribution from the bound states becomes strongly suppressed and hence $\rho_A(E,x)$ goes to zero away from the gated region (compare Fig. \ref{fig4} (c) and Fig. \ref{fig4} (d)). The decay length of the bound states increases with the decrease of ${\left| E \right|}$, and therefore $\rho _B$ has a peak at low energy as shown in Fig. \ref{fig4} (d), which is reflected in the total density of states as shown in Fig. \ref{fig4} (b). Therefore, by comparing $\rho_A(E,x)$ near and away from the gated region using scanning tunneling microscopy, one can identify the contribution from the bound states by $\Delta \rho_A(E)=\rho_A(E,0) - \rho_A(E,-\infty)$. Next, let us consider the tunneling current along the 1D channel, taking into account the electron-electron interaction. Away from the half-filling, the possible scattering processes are dispersive ($g_d$) and forward ($g_f$) scatterings.\cite{wu2006,xu2006} Near the Fermi points, the corresponding interactions can be easily taken into account by the standard bosonization method.\cite{Solyom,Giamarchi,Senechal} The bosonized Hamiltonian is \begin{eqnarray} H = \frac{v}{2}\int {dx\left[ {\frac{1}{K}(\partial _x \phi )^2 + K(\partial _x \theta )^2 } \right]}, \\ v = \sqrt {\left( {v_F \left| {\cos Z} \right| + \frac{{g_f }}{{2\pi }}} \right)^2 - \left( {\frac{{g_d }}{\pi }} \right)^2 }, \\ K = \sqrt {\frac{{2\pi v_F \left| {\cos Z} \right| + g_f - 2g_d }}{{2\pi v_F \left| {\cos Z} \right| + g_f + 2g_d }}} \label{K} \end{eqnarray} with $\phi = \phi _{R \uparrow }^{} + \phi _{L \downarrow }^{}$ and $\theta = \phi _{R \uparrow }^{} - \phi _{L \downarrow }$ where $\phi _{R \uparrow }$ and $\phi _{L \downarrow }$ define chiral boson fields of spin up right mover and spin down left mover, respectively. As seen, the Luttinger parameter $K$ is tunable by changing $Z$, namely the gate voltage. This opens up a possilbility of controlling the property of the 1D interacting fermions by gate voltage. We consider a single magnetic impurity placed on the gated region, which produces the backward scattering. Then, the tunneling current $I_C$ can be calculated within the linear response regime and is given by \cite{Strom,Kane,Furusaki} \begin{eqnarray} I_C \sim \left( {eV} \right)^{\frac{2}{K} - 1} \end{eqnarray} with an applied voltage $V$. Here, we consider a weak impurity scattering. For a strong impurity scattering, the result will be modified due to formation of the localized impurity resonance.\cite{XZhang,Biswas} Thus, the dependence of $K$ on $Z$ can be experimentally studied through the tunneling current. Since the backward scattering requires spin flip process, when the magnetic moment of the magnetic impurity is parallel to the spin of the bound states, the backward scattering is prohibited. This means that by changing the direction of the magnetic moment by, e.g. an applied magnetic field, the tunneling current can be tuned. \begin{figure}[htb] \begin{center} \scalebox{0.8}{ \includegraphics[width=6.5cm,clip]{fig2.eps}} \end{center} \caption{ (a) $K$ and (b) $2/K-1$ as a function of $Z/\pi$ at $g_f/(2\pi v_F)=g_d/(2\pi v_F)=0.1$. } \label{fig2} \end{figure} In Fig. \ref{fig2}, we show (a) $K$ and (b) $2/K-1$ as a function of $Z/\pi$ at $g_f/(2\pi v_F)=g_d/(2\pi v_F)=0.1$. As shown in Fig. \ref{fig2} (a), the $K$ value oscillates with $Z$ as expected from Eq.(\ref{K}). Near $Z=\pi/2$ mod $\pi$, $K$ is strongly suppressed. Correspondingly, the exponent of the current, $2/K-1$, has a diverging behavior near such points as shown in Fig. \ref{fig2} (b). This dependence can be observed in tunneling experiments. It is noted that the conductance between the source and drain attached to the gated region is still dominated by the 2D surface metal on topological insulator, i.e., $\sigma_{2D} = (e^2/h) k_F \ell >> \sigma_{1D} \sim e^2/h$ with the mean free path $\ell$ and the conductance from the 1D channel $\sigma_{1D}$. One possible way to separate these two contributions is the modulation spectroscopy using the precession of the spin of the magnetic impurity. Namely, the backward scattering potential is dependent on the direction of the spin, which oscillates with the precession, and hence the ac component of the 1D conductance occurs. One catch is that the universal conductance fluctuation (UCF) of the order of $e^2/h$ also contributes to the ac component of the conductance \cite{UCF}, but the dependence on the spin direction is random, while the 1D channel conductance shows the systematic dependence, i.e., it is most suppressed when the backward scattering is maximized. Therefore, by taking the average over several samples, one can separate the UCF and the 1D conductance of our interest. In summary, we studied the formation of the one-dimensional channels on the topological surface under the gate electrode. The energy dispersion of these channels is almost linear in the momentum with the velocity sensitively depending on the strength of the gate voltage. The energy is also restricted to be positive or negative depending on the strength of the gate voltage. As a result, the local density of states near the gated region has an asymmetric structure with respect to zero energy. With the electron-electron interaction, the correlation effect in the bound states can be tuned by the gate voltage, which can be verified by tunneling experiments. This work is supported by Grant-in-Aid for Scientific Research (Grants No. 17071007, 17071005, 19048008 19048015, and 21244053) from the Ministry of Education, Culture, Sports, Science and Technology of Japan. T.Y. acknowledges support by JSPS. Work at Los Alamos was supported by US DOE through LDRD and BES (A.V.B).
\subsection{Simplest possible model for membranes and asymmetric diffusion} Managing flows of ions and molecules across membranes is a central task of living systems. Cell membranes contain a wide variety of pores, which can sort, drive, and rectify fluxes of different particles in both directions. There are many models for the controlled leaks, and one-way flows of particular ionic species, which are useful to a cell \cite{hille:01}. The simplest leak is passive diffusion through a pore, driven by a concentration gradient. If we assume noninteracting point particles, the flow can be described by Fick's law. The flux is linear in the concentration gradient across the membrane, the diffusion coefficient is independent of concentration. However, in a biological context, diffusing objects are of finite size, often closely packed, and often not small compared to a pore diameter. These conditions can produce diffusion with a highly nonlinear dependence on concentration, and lead the system to states far from the linear Onsager regime of nonequilibrium thermodynamics. The goal of this paper is to describe a common dynamic arising in such cases, generated solely by kinetic constraints and the geometry of the pore and diffusing objects. A clear departure from linear diffusion takes place when some object enters a pore, and due to size or geometrical orientation cannot pass all the way through. If there are many such blockers on one side of a membrane, carried in a bath of smaller free-flowing particles, we can produce a rectifier. Flow in one direction sweeps blockers into pores, interrupting further flow. Flow in the other direction clears the pores, re-establishing free flow. This mechanism was in fact found in the squid giant axon, by Armstrong and Binstock in 1965 \cite{armstrong:65}. A tacit assumption of many subsequent membrane studies is that there must exist forces which hold blockers in place, so that they are not dislodged by thermal fluctuations. Among proposed mechanisms are electrostatic attraction, due to charges fixed in the pore walls \cite{hille:78}, and microscopic flaps and tethers, to mechanically hold a blocker \cite{armstrong1997commentary,armstrong1977inactivation}. Here we argue that kinetic constraints alone can readily perform this function. A blocker must backtrack to reopen a pore, and as concentration is increased, more and more objects obstruct this motion. Thus pore clearance can become statistically unlikely, requiring a series of favorable fluctuations \cite{Shaw2007geometry,Packard2004asymmetric}. Figure 1 illustrates a number of situations where blocking can arise, diffusing objects may arrive at a configuration that blocks flow through the pore. This is a consequence of simple geometry only, and can occur in models employing Hamiltonian or Monte Carlo dynamics, in continuous space or on a lattice, with two species of diffusing object, or just one. The common theme of all of these models is that flow through a pore enters a blocked configuration, a dead end in the configuration space. In order for the block to be removed, objects must back up a finite distance along the path they travelled. If there is a high concentration of objects in and around a pore, this retracing may be unlikely, and the system will be trapped in a long-lived metastable blocked state. This blocked macroscopic state then becomes a type of attractor, in the sense that most microscopic configurations lead to it. \begin{figure*} \includegraphics[scale=0.5]{fig1} \caption{\label{blocked} Blocking of membrane pores with different pore geometries and different particle types. (a) Large and small disks moving in continuous space. (b) Large and small squares moving on a lattice, with a spacing of the size of the small squares. (c) Dominos moving on a lattice; while in a pore, a domino cannot rotate. (d) Single species of square, moving on a lattice with spacing one-half the size of the square. (e) Single species of disk, in the continuum. In all cases, blockers must move upward to re-establish flow. In model simulations, particles in (a) and (e) move in continuous two-dimensional space, governed by Hamiltonian dynamics. Particles in (b-d) move on a lattice, with Monte Carlo updates. } \end{figure*} Given a high enough concentration of particles, a pore becomes blocked, and remains blocked for a time that depends on concentration and pore geometry, which determine the strength of the attractor. When the stability of the attractor is weak, local density fluctuations around the pore may cause fluctuations between blocked and unblocked pore states, with a consequent intermittent flux transmission through the pore. A hard disk simulation of a membrane consisting of a single pore with the geometry of Fig.~1(a) displays this behavior \cite{PatchClamp}, as shown in Fig.~2. This flux intermittency is qualitatively similar to data obtained from patch clamp observations of current through a single cell membrane pore \cite{hamill1981improved}. \begin{figure}[b] \includegraphics[scale=0.47]{fig2} \caption{\label{mem} A "patch clamp" time series of the flux of small disks through a single pore in the presence of blockers, as in the geometry of Fig.~1(a). The particles are diffusing through the pore from a chamber kept at constant concentration, into a vacuum. The flux values are the number of particles passing through the pore in each interval of 1000 time steps \cite{PatchClamp}. } \end{figure} The interactions between particles and pores that produce the blocking states illustrated in Fig.~1 are simple enough that we might expect this situation to occur quite generally. We now present an illustrative model that uses macroscopic variables to characterize the blocking state of the whole membrane, capturing qualitative behavior common to all the various blocking scenarios. The model membrane contains a large number of pores, each of which can either be open or closed. For open pores, we assume Fickian diffusion and, for simplicity, that one side of the membrane is held at zero concentration. The flux across the membrane is thus $J = fDC$, where $f$ is the fraction of open pores, $D$ is a diffusion coefficient, and $C$ the concentration on the occupied side of the membrane. We describe the probabilities of a single pore entering or leaving the blocked state. The probability that during a small time interval a blocker enters a pore we take to be proportional to concentration, and a coefficient reflecting the fraction of blockers, or blocking configurations, $P_{\rm closes} = B C$. The simulations indicate this assumption to be reasonable. For a blocked pore to become open, the blocker must experience a fluctuation which backs it out of the pore. This will become increasingly unlikely as the concentration behind it increases, and as the pore length increases. If we assume that a series of independent favorable fluctuations are required for the blocker to move backward a finite distance, individual fluctuation probabilities multiply, and we are led in the simplest case to an exponential, $P_{\rm opens} = e^{-GC}$, where $C$ is the concentration difference. $G$ is a geometrical factor that encodes the interaction between the diffusing blocker and the pore, and would depend, for example, on the length of the pore, or the distance the blocker has to retrace. Now we consider a chamber of unit volume, leaking through a membrane with many pores, into a vacuum, and assume the number of pores is large enough so that we can treat the fraction of open pores, $f$, as a continuous variable. We obtain a pair of coupled nonlinear first-order ODE's. \[ {{dC}\over{dt}} = -fDC, \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ {{df}\over{dt}} = (1-f)e^{-GC} - fBC \] The first describes the change of concentration in the chamber due to flux through the available open pores, and the second the change in the fraction of open pores due to the dynamics as described above. If the concentration in the chamber is held constant, we can solve for the flow as a function of concentration in the steady state, obtaining \( J(C) = {DC}/({BCe^{GC}+1}). \) Figure 3 displays the results of a hard-disk effusion simulation \cite{DumpHDisk}, along with curves from the above expression \cite{rectifyHDisk}. The dotted curves are obtained from a Hamiltonian system of two sizes of elastic hard disks moving in a chamber, one wall of which is a membrane perforated with tapered pores with the geometry of Fig. 1(a). The smaller disks are able to effuse through the pores into a vacuum, any that do so are replaced in a random position in the chamber, thus keeping concentrations constant. We examine four different pore lengths. At a given pore, the outward flow may resemble the intermittent behavior of Fig.~2, but we take an average over multiple pores, and over a long time. \begin{figure}[h] \includegraphics[scale=.4]{fig3} \caption{\label{rec} Steady-state flow of small disks through a membrane pore with the geometry of Fig.~\ref{blocked}(a), as a function of pressure in the chamber. The dots are data points from a simulation, and the curves are generated by the steady-state equation above. The relative pore lengths are 2, 3, 5, and 10 (from the top curve down). Flows through the reversed pore geometry are plotted as negative values. } \end{figure} Plotted is the flux through the membrane as a function of the incremental pressure of small disks added to the chamber, with a fixed number of large blocker disks present. This pressure is the appropriate variable governing the dynamical flux in this Hamiltonian system, and is computed from the concentration (disk area) using the hard disk equation of state \cite{Santos:1995p4001}. For low concentrations of small disks, the flow is of Fick's law form, with an effective diffusion constant D. But as the concentration is increased, large disk blockers are swept into pores, and the mean flow is decreased. At high concentrations, the probability that a blocker can back out of the pore becomes small, and the flow is reduced to near zero. If the membrane geometry is reversed, that is, disks strike the membrane from the flat side, or bottom-up in Fig. 1(a), the large disks are not trapped in the pore by trailing small disks. In this case, the flux remains a linear function of pressure up to high concentrations, these values are plotted as negative values in Fig.~3. Note that the asymmetric geometry of the membrane leads directly to asymmetric diffusion in the two flow directions. The solid curves of Fig.~3 are plotted from the analytic expression, we take G values in proportion to the simulation pore lengths, and fit the theoretical curves to the simulation data with a single overall scaling. The blocking factor B varies from 1, when all disks present in the chamber are large-disk blockers, to smaller values, as an increasing fraction of small disks are added \cite{rectifyHDisk}. The diffusion constant D is found from the simulation data, in the linear low concentration range. The fit is quite satisfactory, considering the pores in the simulation are actually conical, and that density fluctuations of the hard disk gas are nontrivial \cite{Salacuse:2007p4002}. In fact it will be more difficult for a blocker to escape than our exponential estimate predicts, because a fluctuation which empties a pore so that a blocker can back out will create a locally larger concentration near the mouth of the pore. For extended particles, such a fluctuation will be reflected in a very large local pressure increase, and will be highly unlikely as concentration increases \cite{Santos:1995p4001,Salacuse:2007p4002}. The model equations will not be particularly accurate for all pore and blocker geometries. But they may be a minimal set, describing deviations from the linear regime of nonequilibrium flow. All of the geometries in Fig.~\ref{blocked} generate qualitatively similar simulation curves. This model has the form of a biased random walk, on a line with a reflecting barrier on one end (corresponding to the blocked end of the pore) and an absorbing barrier on the other end (the open end of the pore where the particle is absorbed by the bulk). The bias is due to the mean flow. Previous discussions of diffusion through pores have used similar, though unbiased, Brownian walk models. However the physical interpretations of the random walk are completely different, for example describing a blocker tethered outside a pore \cite{millhauser1988diffusion}, or conformational transformations of the pore itself \cite{Goychuk:2002p1912}. In our simple model, blockers are biased into a pore by a concentration difference, that is, a chemical potential. We would like to note that a model of identical form is produced by considering a single charged blocker moving in a pore, biased by an electrical potential maintained across the membrane. This might be a minimal model for a voltage-controlled rectifier, an important functional component of many cell membranes \cite{hille:01}. Other models of voltage control feature charged sensor levers, and other mechanical degrees of freedom which must move to restrict ion flow. The model we present here can produce voltage control with no additional degrees of freedom, just a rigid pore with appropriate geometry. Finally we examine the behavior of a leaky chamber filled initially at a high concentration, and then allowed to decline in time. In Fig.~4 we compare the model equations with a simulation of the domino geometry of Fig.~1(c) \cite{dominoRules}, though any of the models of Fig.~1 produce qualitatively similar results. While the concentration gradient across the membrane is large, the leak rate is small. But eventually the seal loosens, and there is a relatively sudden dumping of the contents of the chamber. For the domino simulations, note the large effect of fluctuations. Even in this simulation, with a chamber initially containing 35,000 dominos, and a membrane of length 4000 with 500 pores, the onset of the high flux region is quite variable. In both the simulation and the model equations, there is visible a short transient to an attracting state where most pores are blocked, from an initial condition where all pores are open. \begin{figure}[h] \includegraphics[scale=0.6]{fig4} \caption{\label{dump} Simulation of the concentration of a leaky chamber of dominos, as a function of time. Shown at the top are six runs with the same initial concentration, for two pore lengths with the geometry of Fig.~1(c). Below are curves generated by the model coupled equations, for a few values of G. The flux associated with the G=8 curve is plotted as a dotted line. } \end{figure} Clearly, careful studies of the experimental consequences of this model are required. For example, the statistical properties of curves like that of Fig.~2 should be compared with real patch clamp data. Another important issue is temperature dependence. The escape probability of a blocker held by a kinetic constraint would not be expected to have the Arrhenius temperature dependence of a blocker held in a potential well. Our discussion has emphasized the role of kinetic constraints and geometry in influencing flows of extended objects, via the simple requirement that a system back up a finite distance in its configuration space before proceeding further. This picture of dead-end alleys in configuration space may be of some general use in modeling other nonequilibium systems. However, geometry alone cannot be the whole story. A real membrane pore carries charged ions, and is constructed of proteins with charged residues; electric fields must play a central part. A model including electric charges fixed to pore walls explains observed asymmetric diffusion in fabricated conical nanopores \cite{Siwy:2004p567}. Also, living cell membranes may have evolved a number of intricate mechanisms for controlling ionic flow. Nevertheless we suggest that, due to its simplicity, the blocking mechanism described here will often be present, and moreover it may have been the dominant mechanism for simpler pores that occurred early in evolutionary history. R. Shaw gratefully acknowledges the hospitality of Theo Geisel at the Max Planck Institute for Dynamics and Self-Organization, and financial support by the German Federal Ministry for Education and Research (BMBF) via the Bernstein Center for Computational Neuroscience (BCCN) G\"ottingen under Grant No. 01GQ0430.
\section{Introduction} Spintronics aims at using the information carried by the electrons' spin in solids. For this purpose, establishing reliable methods to create, transfer and detect the spin current is an urgent task. Compared to the charge transports, spin transports have one serious fundamental difficulty. That is the non-conservation of the spin in solids. This limits the range of the spin transmission to be less than the spin diffusion length, which is typically $\mu$m scale in metals. The non-conservation of the spins is expressed by a source term in the continuity equation for the spin \begin{align} \dot{s}^\alpha+\nabla \cdot\bm{j}_{\rm s}^\alpha={\cal T}^\alpha. \label{sconteq} \end{align} Here $s$ and $\bm{j}_{\rm s}$ are the spin density and spin current, respectively, $\alpha=x,y,z$ is the spin direction and ${\cal T}$ is the spin relaxation torque resulting in the non-conservation of the spin. In the most cases in metals, the dominant origin of ${\cal T}$ is the spin-orbit interaction. Although the relaxation torque term is essential in spin transports, it has so far been treated only on the phenomenological ground. The continuity equation is equivalent to the Boltzmann equation, which is useful in discussing the spin transports. The Boltzmann equation for the distribution function of each spin channel was discussed by Son et al. \cite{Son87} and later by Valet and Fert \cite{Valet93} in the context of the giant magnetoresistance in multilayer systems. In their analysis, they approximated the spin relaxation torque as proportional to the inverse of a spin relaxation time $\tau_{\rm sf}$ and to some unknown function representing a driving force for the spin accumulation. The driving force was written in terms of what they called the spin chemical potential $\mu_{\rm s}$. The relaxation torque was approximated as ${\cal T} ^z =\mu_{\rm s}/\tau_{\rm sf}$. They argued that $\mu_{\rm s}$ satisfies the diffusion equation, $\nabla^2 \mu_{\rm s} =-\ell_{\rm sf}^{-2} \mu_{\rm s}$, with the diffusion length $\ell_{\rm sf}\propto \sqrt{\tau_{\rm sf}}$. Microscopic calculation for $\mu_{\rm s}$ has not been done so far. The diffusion equation for the spin has been widely used to discuss recent spin transports in metallic junctions \cite{Takahashi08}. The decay of spin transport has been confirmed in non-local spin injection experiments \cite{Jedema01,Kimura07,Seki08}, which indicate that the spin diffusion decays with a decay length of 350-500nm in Cu and 100nm in Au at room temperature. Although the spin diffusion equation appears to be so far successful, the phenomenological treatment of the spin relaxation term and the spin chemical potential must be improved to consider the spin transport seriously. Besides diffusive spin current, there is another spin current that is driven by an effective field. In contrast to the diffusive one, this field-driven contribution should not decay in uniform (single domain) ferromagnets, since the ratio of the spin current and the charge current is determined by the spin polarization ratio of the material, which is a statistical mechanical quantity. The field-driven (local) spin current and the diffusive spin current behave differently, as was recently demonstrated theoretically in the case of the inverse spin Hall effect \cite{Takeuchi10}. In the field of the current-driven magnetization dynamics, the spin relaxation torque has been studied from the microscopic viewpoint \cite{KTS06,TKS_PR08,TE08}. In this context, Eq. (\ref{sconteq}) gives the expression for the torque acting on the spin density ${{\bm s}}$ as \begin{align} \tau^\alpha = -\nabla \cdot\bm{j}_{\rm s}^\alpha+{\cal T}^\alpha. \end{align} In the adiabatic limit, i.e., slowly varying magnetization, and under uniform current, the first term reduces to $\nabla \cdot\bm{j}_{\rm s}^\alpha=(P/2e)(\bm{j}\cdot\nabla){{\bm s}}^\alpha$, where $P$ is the spin polarization of the current \cite{TKS_PR08,TE08}, namely to the adiabatic spin-transfer torque. When the spin-relaxation sets in, the conduction electron no longer follows the magnetization profile, and new contribution to the torque arises from the ${\cal T}$ term. This torque was shown to be \begin{align} {\cal T} = -\beta \frac{P}{es^2} ({{\bm s}}\times (\bm{j}\cdot\nabla){{\bm s}}), \label{betatorque} \end{align} where $\beta$ is a coefficient inversely proportional to the spin relaxation time $\tau_{\rm s}$ \cite{KTS06,TE08}. This torque, called $\beta$ term, turned out to be essential in determining the efficiency of the current-driven domain wall motion \cite{TK04,Zhang04,Thiaville05}. The magnitude of the parameter $\beta$ has recently been intensively studied experimentally by measuring the domain wall speed under current \cite{Thomas06,Moore09}. Theoretical formulation for estimating $\beta$ in the first-principles calculations was carried out recently \cite{Garate09}. The spin relaxation torque has been studied also from the viewpoint of how to define the spin current. It was discussed that the spin relaxation torque contains a term written as a divergence of the torque dipole density, ${\bm P}$ \cite{Culcer04}. Generalized argument was given by Shi et al. \cite{Shi06}, where they discussed that the $z$ component of the relaxation torque is written as a divergence, \begin{align} {\cal T}^z= -\nabla\cdot{\bm P}, \end{align} if the system has the inversion symmetry. This means that the total torque integrated over the system should vanish. Shi et al. also argued that if the relaxation torque is a divergence of ${\bm P}$, one can define a spin current that is conserved. In fact, defining $\tilde{\bm{j}_{\rm s}}\equiv \bm{j}_{\rm s}+{\bm P}$, the continuity equation (\ref{sconteq}) reduces to $\dot{s}^z+\nabla\cdot \tilde{\bm{j}_{\rm s}}=0$. The explicit form of the torque dipole density was not calculated in Ref. \cite{Shi06}. Obviously, in the presence of the inhomogeneity of the magnetization, the $\beta$ torque (\Eqref{betatorque}) cannot be written as a divergence, and thus it indeed represents the spin angular momentum lost by the spin relaxation. The result that the spin relaxation torque is given by a derivative of the applied electric field ${\bm E}$ is understood as follows. The spin relaxation torque should of course vanish when ${\bm E}=0$. It cannot be directly proportional to ${\bm E}$, since field-driven spin current in uniform ferromagnets should not decay. Therefore, the simplest expression for the relaxation torque is a derivative of the field. It is not, however, obvious whether it should be always written in a rotationally invariant way, or if it can be anisotropic, since the rotational invariance is broken in uniform ferromagnets because of the magnetization. The first aim of the present paper is to calculate the spin relaxation torque microscopically in the presence of the applied electric field. The spin relaxation mechanism we take into account is the spin-orbit interaction due to random impurities. Our explicit calculation reveals that the spin relaxation torque is not always rotationally symmetric in uniform ferromagnets, but is generally given by \begin{align} {\cal T}^z = \gamma (\nabla\cdot {\bm E}) +\delta\gamma(\partial_z E_z), \label{TauE} \end{align} where $z$ axis is along the magnetization, $\gamma$ and $\delta\gamma$ are coefficient proportional to the inverse spin relaxation time. The spin relaxation torque is, therefore, anisotropic. Relaxation torque of \Eqref{TauE} indicates that the torque dipole density is given by \begin{align} {\bm P} = -(\gamma {\bm E} +\delta\gamma ({\bm n}\cdot {\bm E}){\bm n}), \end{align} where ${\bm n}$ represents the direction of the magnetization. These anisotropic behaviors of the transport quantities is common when spin-orbit interaction exists, as is well-known in charge transport as the anisotropic magnetoresistance (AMR) \cite{Mcguire75}. The second aim of the paper is to study the spin current on the same microscopic footing as the relaxation torque. We show that the spin current is made up of a field-driven contribution which is local and the diffusive one with nonlocality. The field-driven contribution is anisotropic like the AMR effect for the charge current (we call the effect as spin AMR effect). The diffusive contribution is given as a gradient of a spin chemical potential, $\mu_{\rm s}$. We will derive the linear-response expression for the spin chemical potential. Our microscopic study on the spin current demonstrates the validity of the half-phenomenological treatments \cite{Valet93}. \section{Model} We consider the conduction electron system taking account of the spin-orbit interaction, the impurity scattering without spin flip, and the applied electric field. The Hamiltonian of the system is given as $H=H_0+ {H_{\rm so}}+H_{\rm em}+H_{\rm imp}$, where $H_0$ is the free electron Hamiltonian including the uniform magnetization, ${H_{\rm so}}$ is the spin-orbit interaction, $H_{\rm em}$ is the interaction with the gauge field representing the applied electric field, and $H_{\rm imp}$ is the spin-independent impurity scattering. The free part reads \begin{align} H_0\equiv \sum_{{\bm k}\sigma}\epsilon_{{\bm k}\sigma} c^\dagger_{{\bm k}\sigma} c_{{\bm k}\sigma} , \end{align} where the electron creation and annihilation operators are denoted by ${c^{\dagger}}$ and $c$, respectively, $\epsilon_{{\bm k}\sigma}\equiv \frac{k^2}{2m}-{\epsilon_F}-\sigma {M}$, ${\epsilon_F}$ is the Fermi energy, ${M}$ is the spin splitting due to the magnetization and $\sigma\equiv \pm$ represents the spin. The spin-orbit interaction is represented by ${H_{\rm so}}={H_{\rm so}}^0+{H_{\rm so}}^{A}$, where ($\sigma_k$ ($k=x,y,z$) is the Pauli matrix) \begin{align} {H_{\rm so}}^0 &= -\frac{i}{2} \sum_{ijk} \epsilon_{ijk}\int\! {d^3x} (\nabla_i v_{\rm so}^{(k)}) ({c^{\dagger}} \vvec{\nabla}_j \sigma_k c) \\ {H_{\rm so}}^{A} &= - e\sum_{ijk} \epsilon_{ijk}\int\! {d^3x} (\nabla_i v_{\rm so}^{(k)}) A_j({\bm x},t) ({c^{\dagger}} \sigma_k c). \end{align} (We suppress the spin index when obvious, namely, $c=(c_+,c_-)$. ) The spin-orbit potential $v_{\rm so}^{(k)}$ is assumed to arise from random impurities and to depend on the spin direction ($k$). The averaging over the spin-orbit potential is carried out as \begin{align} \average{ v_{\rm so}^{(k)}({\bm p}) v_{\rm so}^{(\gamma)}(-{\bm p}')}_{\rm i} =n_{{\rm so}}{\lambda_{\rm so}}^2 \delta_{{\bm p},{\bm p}'}\delta_{k\gamma}, \label{vsav} \end{align} where $n_{{\rm so}}$ and ${\lambda_{\rm so}}$ are the concentration of the spin-orbit impurities and the strength of the interaction, respectively. The average of the spin-orbit potential at the linear order is zero in our model, and thus we do not take account of the the anomalous Hall and spin Hall effects. We consider a case where electric field ${\bm E}({\bm x},t)(\equiv -\dot{{\bm A}}({\bm x},t))$ is position and time dependent. The electromagnetic interaction is written as \begin{align} H_{\rm em} = -\frac{e}{m} \sum_{{\bm k},{\bm q}} \sum_i k_i A_i({\bm q},\Omega) ({c^{\dagger}}_{{\kv}-\frac{{\bm q}}{2}}c_{{\kv}+\frac{{\bm q}}{2}}), \end{align} where $\Omega$ is the frequency of the electric field. We will consider the limit of small $\Omega$ and small $q$. The scattering by the normal impurities is represented by \begin{align} H_{\rm imp} &= \sum_{i=1}^{N_{\rm imp}}\sum_{{\bm k}\kv'} \frac{v_{\rm imp}}{N} e^{i({\bm k}-{\bm k}')\cdot{\bm R}_i} {c^{\dagger}}_{{\bm k}'}c_{{\bm k}}, \end{align} where $v_{\rm imp}$ represents the strength of the impurity potential, ${\bm R}_i$ represents the position of random impurities, $N_{\rm imp}$ is the number of impurities, and $N\equiv V/a^3$ is number of sites. To estimate physical quantities, we take the random average over impurity positions in a standard manner \cite{TKS_PR08}. To derive the spin continuity equation, \Eqref{sconteq}, we derive the equation of motion for the spin density, ${\bm \se}({\bm x},t)\equiv \average{{c^{\dagger}}({\bm x},t){\bm \sigma} c({\bm x},t)}$ ($\average{\ }$ represents the quantum average). The time development of the spin density reads \begin{align} \dot{s}^\alpha &= i\average { [H,c^\dagger]\sigma^\alpha c+ c^\dagger\sigma^\alpha [H,c] },\label{eqcommutators} \end{align} where $H$ is the total Hamiltonian of the system. The commutators are calculated as in Appendix \ref{APP:eqofmo}, and \Eqref{eqcommutators} turns out to be \Eqref{sconteq}, namely \begin{align} \dot{s}^\alpha = -\nabla \cdot\bm{j}_{\rm s}^\alpha+{\cal T}^\alpha, \nonumber \end{align} with the spin current given as \begin{align} j_{\rm s}^{\alpha} \equiv j_{\rm s}^{{\rm (n)},\alpha}+j_{\rm s}^{{\rm so},\alpha} , \end{align} where \begin{align} \jspin{,i}{{\rm (n)},\alpha} &\equiv -\frac{i}{2m} \average{ {c^{\dagger}} \sigma^\alpha \vvec{\nabla}_i c } -\frac{e}{m}A_i \average{ {c^{\dagger}} \sigma_\alpha c } \nonumber\\ &\equiv \jspin{,i}{(0),\alpha} + \jspin{,i}{A,\alpha} , \label{spincurrentsdefs} \end{align} and \begin{align} \jspin{,i}{{\rm so}, \alpha} \equiv -\sum_{j} \epsilon_{ij\alpha} (\nabla_jv_{\rm so}^{(\alpha)}) \average{ {c^{\dagger}} c}. \label{jssodef} \end{align} The relaxation torque reads \begin{align} {\cal T}^{\alpha} \equiv {\cal T}_{{\rm so}}^{\alpha} +{\cal T}_{{\rm so}}^{A,\alpha}, \label{totaltaudef} \end{align} where \begin{align} {\cal T}_{{\rm so}}^{\alpha} &\equiv i \sum_{ijkl} \epsilon_{ijk} \epsilon_{\alpha lk} (\nabla_iv_{\rm so}^{(k)}) \average{ {c^{\dagger}} \sigma_l \vvec{\nabla}_j c}, \label{Tsodef} \\ {\cal T}_{{\rm so}}^{A,\alpha} & \equiv 2e \sum_{ijkl} \epsilon_{ijk} \epsilon_{\alpha lk} (\nabla_iv_{\rm so} ^{(k)} ) A_j \average{ {c^{\dagger}} \sigma_l c}.\label{TsoAdef} \end{align} The spin relaxation torque depends on the definition of the spin current. For instance, if we redefine the spin current as ${\bm{j}_{\rm s}' } ^\alpha \equiv {\bm{j}_{\rm s}^\alpha }-{\bm C}^\alpha$, where ${\bm C}$ is a vector, the continuity equation (\ref {sconteq}) becomes $\dot{s}^\alpha+\nabla \cdot{ \bm{j}_{\rm s}' } ^\alpha={{\cal T}'} ^\alpha$, where the relaxation torque reads ${{\cal T}'} ^\alpha\equiv {{\cal T}} ^\alpha+\nabla\cdot {\bm C}^\alpha$. This ambiguity of spin current definition of course does not affect physical quantities such as the total torque acting on the spin density, which is given by $\dot{{{\bm s}}}$. \section{Spin relaxation torque} We calculate the spin relaxation torque as a linear response to the applied electric field. The uniform magnetization is chosen as along $z$ axis. The spin-orbit interaction is included to the second order. \begin{figure}[tbh] \begin{center} \includegraphics[width=0.45\hsize]{spinrelaxationdiag6.eps} \caption{ Feynman diagrams representing the relaxation torque. Solid lines represent the electron Green's function with the lifetime ($\tau_\sigma$) included, $v_{\rm so}$ represents the spin-orbit interaction (double dashed line) and dotted line represents the interaction with the gauge field ($A$). The first two diagrams are the contributions to ${\cal T}_{{\rm so}}^{\alpha}$ and the last diagram is the contribution to ${\cal T}_{{\rm so}}^{A,\alpha}$. The vertex marked by cross represents the relaxation torque. \label{FIGspinrelxationdiag1} } \end{center} \end{figure} The contributions to the relaxation torque, \Eqref{totaltaudef}, are shown in Fig. \ref{FIGspinrelxationdiag1}. The leading contribution for small $1/({\epsilon_F}\tau)$ and $q\ell$ ($\ell$ is the electron mean free path) turns out to be the first diagram in Fig. \ref{FIGspinrelxationdiag1}, which reads (see Appendix \ref{SEC:appA} for details) \begin{align} {\cal T}^{\alpha} &= \delta_{\alpha,z} \frac{2}{45\pi} n_{{\rm so}} {\lambda_{\rm so}}^2 \frac{e}{m^2} (3\nabla\cdot\dot{{\bm A}}+\nabla_z\dot{A}_z) \nonumber\\ &\times \sum_{{\bm k}\kv'} k^2 (k')^4 \sum_{\sigma=\pm} \sigma \green{{\bm k},\sigma}{{\rm r}} \green{{\bm k}',-\sigma}{{\rm r}} (\green{{\bm k}',-\sigma}{{\rm a}})^2 +{\rm c.c.}, \label{Tauzdominant} \end{align} where $\green{{\bm k}\sigma}{{\rm r}}$ and $\green{{\bm k}\sigma}{{\rm a}}$ are the retarded and advanced electron Green's functions, respectively, carrying the wave vector ${\bm k}$ and spin $\sigma$ with zero frequency. As we see, only $z$ component of the torque is finite. The Green's functions include the lifetime arising from the self-energy process due to normal impurities and the spin-orbit interaction. The inverse lifetime for the electron with spin $\sigma(=\pm)$ is given as \begin{align} {\tau_\sigma}^{-1} = 2\pi n_{\rm imp}v_{\rm imp}^2 \nu_\sigma (1+\kappa_{z,\sigma}+\kappa_{\perp} \gamma_\sigma), \end{align} where $\nu_\sigma$ is the spin-resolved electron density of states, $n_{\rm imp}$ and $v_{\rm imp}$ are the concentration and the potential strength of the impurities, $\kappa_{z,\sigma}\equiv \frac{1}{3}\frac{n_{{\rm so}}{\lambda_{\rm so}}^2}{n_{\rm imp}v_{\rm imp}^2}k_{F\sigma}^4 $ and $\kappa_{\perp} \equiv \frac{2}{3}\frac{n_{{\rm so}}{\lambda_{\rm so}}^2}{n_{\rm imp}v_{\rm imp}^2}k_{F+}^2 k_{F-}^2 $ are dimensionless ratios of the spin-orbit interaction to the normal impurity scattering ($k_{F\sigma}$ is the spin-dependent Fermi wavelength), and $\gamma_\sigma\equiv\frac{\nu_{-\sigma}}{\nu_\sigma}$. The total relaxation torque is therefore given by (\Eqref{TauE}), \begin{align} {\cal T}^z = \gamma (\nabla\cdot {\bm E}) +\delta\gamma(\partial_z E_z), \end{align} where \begin{align} \gamma &\equiv \frac{8\pi e}{15m^2} n_{{\rm so}} {\lambda_{\rm so}}^2 \dos_{+} \dos_{-} k_{F+}^2 k_{F-}^2 \sum_{\sigma=\pm} \sigma k_{F\sigma}^2\tau_{\sigma}^2 \end{align} and $\delta\gamma\equiv \gamma/3$. The parameter $\gamma$ is proportional to the spin flip rate due to the spin-orbit interaction. Our result indicates that the relaxation torque is zero in uniform ferromagnet when uniform electric field is applied. Thus the spin current does not decay in this case. In fact, the static solution of \Eqref{sconteq} with ${\cal T}^z=0$ is $j_{\rm s}^z={\rm constant}$. The degree of the asymmetry, $\delta\gamma/\gamma$, is not universal but is model dependent. For instance, in the case of junction with weak electron hopping at point-like leads, $\delta \gamma$ vanishes \cite{Takezoe10}. \section{Spin current} We here calculate the spin current within the same formalism. Within the linear response theory, the spin current is calculated by estimating the Feynman diagrams shown in Figs. \ref{FIGspincurrent} and \ref{FIGdiffusion}, which correspond to the field-induced contribution and the effect of the diffusive electron motion, respectively. \begin{figure}[tbh] \begin{center} \includegraphics[width=0.6\hsize]{spincurrentdiag6.eps} \caption{ Feynman diagrams representing the local contribution to the spin current (the first two terms in \Eqref{jsexpression2}). The first three diagrams correspond to $ \jspin{}{{\rm (n)}}$, and the last two diagrams represent the contribution from the anomalous velocity, $\jspin{}{{\rm so}}$. Dotted and double dashed lines denote the interaction with the applied electric field and the spin-orbit interaction, respectively. The vertex marked by cross represents the spin current. \label{FIGspincurrent} } \end{center} \end{figure} \begin{figure}[tbh] \begin{center} \includegraphics[width=0.65\hsize]{spincurrentdiffusive3.eps} \caption{ Left: The vertex correction contribution to $j_{\rm s}^{{\rm (n)},z}$, resulting in the diffusive spin current (the last term of \Eqref{jsexpression2}). Right: $\Gamma_{\sigma\sigma'}$, which is a ladder process of the successive electron scattering by the normal impurity and the spin-orbit interaction (represented by thick dotted lines) connecting the spin indices $\sigma$ and $\sigma'$. \label{FIGdiffusion} } \end{center} \end{figure} We first estimate the normal part of the spin current, $\jspin{,i}{{\rm (n)},z}$ (\Eqref{spincurrentsdefs}), shown in the first two diagrams in Fig. \ref{FIGspincurrent}. The contribution $\jspin{,i}{{\rm (n)},z}$ is defined including the anomalous contribution from the electromagnetic gauge field, $\jspin{,i}{A,z}$. Its dominant contribution in the limit of small $\Omega$ and small $q$ is calculated as (see Appendix \ref{SEC:appB} for detail) \begin{align} \jspin{,i}{{\rm (n)},z} &= -i\frac{1}{2\pi} \frac{e^2}{m^2} \sum_{{\bm k}{\bm q}}e^{-i{\bm q}\cdot{\bm x}}\int\!\frac{d\Omega}{2\pi} e^{i\Omega t} \sum_{j} \Omega A_j({\bm q},\Omega) \sum_{\sigma} \sigma \left[ k_i k_j\green{{\kv}-\frac{{\bm q}}{2},\sigma}{{\rm r}} \green{{\kv}+\frac{{\bm q}}{2},\sigma}{{\rm a}} \right. \nonumber\\ & \left. + \sum_{{\bm k}'\sigma'} k_i k'_j \green{{\kv}-\frac{{\bm q}}{2},\sigma}{{\rm r}} \green{{\kv}+\frac{{\bm q}}{2},\sigma}{{\rm a}} \green{{\kvp}-\frac{{\bm q}}{2},\sigma'}{{\rm r}} \green{{\kvp}+\frac{{\bm q}}{2},\sigma'}{{\rm a}} n_{\rm imp}v_{\rm imp}^2 \Gamma_{\sigma\sigma'}({\bm q},\Omega) \right] \label{spincurrent1} \end{align} In \Eqref{spincurrent1}, the first term is the contribution shown in the left of Fig. \ref{FIGspincurrent}, and the second term is the contribution from the vertex correction (Fig. \ref{FIGdiffusion}). The factor $\Gamma_{\sigma\sigma'}({\bm q},\Omega)$ contains all the vertex corrections due to the normal impurities and the spin-orbit interaction shown in Fig. \ref{FIGdiffusion}. The equation of motion for $\Gamma_{\sigma\sigma'}$ is derived in the same manner as in Ref. \cite{Hikami80} carried out in the context of quantum correction (the diffusion without the spin-orbit interaction was considered in Ref. \cite{Ban09}). The equation is obtained as \begin{align} \Gamma_{\sigma\sigma} &= (1+\kappa_z)(1+\Pi_\sigma\Gamma_{\sigma\sigma}) +\kappa_\perp \Pi_{-\sigma}\Gamma_{-\sigma,\sigma} \nonumber\\ \Gamma_{\sigma,-\sigma} &= \kappa_\perp(1+\Pi_{-\sigma}\Gamma_{-\sigma,-\sigma}) +(1+\kappa_z) \Pi_{\sigma}\Gamma_{\sigma,-\sigma}, \label{Gammaeqs} \end{align} where \begin{align} \Pi_\sigma({\bm q},\Omega) &\equiv n_{\rm imp}v_{\rm imp}^2 {\sum_{\kv}} \green{{\kv}-\frac{{\bm q}}{2},\sigma}{{\rm r}} \green{{\kv}+\frac{{\bm q}}{2},\sigma}{{\rm a}} \nonumber\\ &\simeq [1-(D_\sigma q^2 \tau_\sigma+\kappa_z+\kappa_\perp\gamma_\sigma)], \label{Pires} \end{align} where $D_\sigma \equiv \frac{(k_{F\sigma})^2}{3m^2}\tau_\sigma$ is the diffusion constant. Neglecting quantities of order of $(\kappa_z,\kappa_\perp)^2$, \Eqref{Gammaeqs} is solved as \begin{align} \Gamma_{\sigma\sigma} &= \frac{1+\kappa_z-(1+2\kappa_z)\Pi_{-\sigma}} {[1-(1+\kappa_z)\Pi_+][1-(1+\kappa_z)\Pi_-]} \nonumber\\ \Gamma_{\sigma,-\sigma} &= \frac{\kappa_\perp} {[1-(1+\kappa_z)\Pi_+][1-(1+\kappa_z)\Pi_-]}. \label{Gammaeqs2} \end{align} Using \Eqref{Pires}, we obtain $\Gamma_{\sigma\sigma'}$ as (assuming the rotational symmetry for the wave vectors when averaging over the spin-orbit potential) \begin{align} \Gamma_{\sigma\sigma} &= \frac{1} {D_\sigma q^2 \tau_\sigma+\kappa_\perp\gamma_\sigma} \nonumber\\ \Gamma_{\sigma,-\sigma} &= \frac{\kappa_\perp} {[D_+ q^2 \tau_++\kappa_\perp\gamma_+] [D_- q^2 \tau_-+\kappa_\perp\gamma_-]}. \label{Gammaeqs3} \end{align} By use of \Eqref{Gammaeqs3} and summing over the wave vectors, the normal spin current, \Eqref{spincurrent1}, reads \begin{align} \jspin{,i}{{\rm (n)},z} &= \sigma_{{\rm s}}^0 E_i - \nabla_i \mu_{\rm s}, \label{jsexpression0} \end{align} where $\sigma_{{\rm s}}^0 \equiv e \sum_{\pm} (\pm) D_\pm \nu_\pm$ is the bare spin conductivity divided by $e$. The first term of \Eqref{jsexpression0} is the field-driven contribution. The second gradient term is a diffusive contribution (vertex corrections), arising from the spin accumulation. The effective potential describing the spin accumulation, $\mu_{\rm s}$, reads \begin{align} \mu_{\rm s} \equiv \int\! {d^3x}' \chi({\bm x}-{\bm x}') (\nabla\cdot {\bm E})({\bm x}'), \label{muspin} \end{align} where $\chi$ is a correlation function arising from the electron diffusion, given as ($V$ is the system volume) \begin{align} \chi({\bm x}) &\equiv - \sum_{\pm} (\pm) \sigma_{\pm} \frac{1}{V}{\sum_{\qv}} \frac{e^{-i{\bm q}\cdot{\bm x}}} {q^2+(\ls{,\pm})^{-2}}. \label{chidef} \end{align} Here $\sigma_\pm \equiv e D_\pm \nu_\pm$ is the spin-resolved Boltzmann conductivity divided by $e$, and the correlation length is given as \begin{align} \ls{,\sigma} = \sqrt{D_\sigma\tau_{{\rm s},\sigma}}. \end{align} The lifetime of the spin $\sigma$ electron reads $\tau_{{\rm s},\sigma} \equiv {\tau_\sigma}/{(\kappa_{\perp} \gamma_\sigma)}$. Defining $\mu_{\rm s}=\mu_+-\mu_-$, we see that spin-resolved effective potential satisfies \begin{align} (-\nabla^2 + (\ls{,\sigma})^{-2}) \mu_\sigma = -\sigma_\sigma (\nabla\cdot{\bm E}) .\label{mudiffusioneq} \end{align} In three-dimensions, the correlation function reads \begin{align} \chi({\bm x})=\frac{1}{4\pi |{\bm x}|}\sum_{\pm} (\pm){\sigma_{\pm}}e^{-|{\bm x}|/\ls{,\pm}}. \end{align} The local part of the spin current arises also from the anomalous spin current due to the spin-orbit interaction, defined in \Eqref{jssodef}. This contribution is calculated by evaluating the last two diagrams in Fig. \ref{FIGspincurrent} as (see Appendix \ref{SEC:appB}) \begin{align} \jspin{,i}{{\rm so},z} &= \delta \sigma_{{\rm s}} (1-\delta_{i,z}) E_i, \label{jssores2} \end{align} where \begin{align} \delta \sigma_{{\rm s}} &\equiv \frac{\pi}{9} n_{{\rm so}}{\lambda_{\rm so}}^2 \frac{e}{m} \sum_{\pm}(\pm) (k_{F\pm})^4 (\nu_{\pm})^2 \tau_\pm . \label{jssores3} \end{align} This spin-orbit correction to the spin conductivity is anisotropic, resulting in a spin version of the anisotropic magnetoresistance (AMR) effect, namely, spin AMR effect. From Eqs. (\ref{spincurrent1})(\ref{Gammaeqs3})(\ref{jssores2}), the leading contribution to the spin current for small ${\bm q}$ and $\Omega$ is obtained as the sum of the local part driven by the electric field and the diffusive part as \begin{align} \bm{j}_{{\rm s}}^{z}=\sigma_{{\rm s}} {\bm E} -\delta\sigma_{\rm s} ({\bm n}\cdot{\bm E}) {\bm n} - \nabla \mu_{\rm s} , \label{jsexpression2} \end{align} where $\sigma_{{\rm s}} \equiv \sigma_{{\rm s}}^{0} +\delta \sigma_{{\rm s}} $ and ${\bm n}$ is the unit vector along the magnetization. In terms of the angle $\theta$ defined by $\cos\theta\equiv ({\bm n}\cdot{\bm E})/E$, the magnitude of the field-driven (local) current reads \begin{align} j_{{\rm s}}^{{\rm loc},z}=\sqrt{ (\sigma_{{\rm s}\parallel})^2 + ((\sigma_{{\rm s}\perp })^2-(\sigma_{{\rm s}\parallel})^2) \sin^2\theta}, \label{jsmag} \end{align} where $\sigma_{{\rm s}\parallel}\equiv \sigma_{{\rm s}}^{0}$ and $\sigma_{{\rm s}\perp}\equiv \sigma_{{\rm s}}$. When the degree of the anisotropy is small, the spin current becomes \begin{align} \frac{j_{{\rm s}}^{{\rm loc},z}}{E} =\sigma_{{\rm s}\parallel}\left(1+ \frac{1}{2}\left( \frac{\sigma_{{\rm s}\perp }}{\sigma_{{\rm s}\parallel}} \right)^2 \sin^2\theta \right). \label{jsmag2} \end{align} We define the magnitude of the spin AMR as \begin{align} \frac{\Delta\rho_{{\rm s}}}{\rho_{{\rm s}\perp}} & \equiv \frac{\rho_{{\rm s}\parallel}-\rho_{{\rm s}\perp}} {\rho_{{\rm s}\perp}} = \frac{ \delta \sigma_{{\rm s}} / \sigma_{{\rm s}} }{1-\delta \sigma_{{\rm s}} / \sigma_{{\rm s}}}, \end{align} where $\rho_{{\rm s}\alpha} \equiv (\sigma_{{\rm s}\alpha})^{-1}$ ($\alpha=\parallel,\perp$). \section{Spin injection} We have thus derived the explicit expression for the spin chemical potential within the linear response theory. Let us apply \Eqref{muspin} to a ferromagnetic-normal metal junction with an insulating barrier, used in the nonlocal spin injection experiments \cite{Kimura07}, depicted in Fig. \ref{FIGspininjection}(a). When the voltage is applied perpendicular to the interface (we choose the $x$ axis in this direction), the electric field is uniform inside the ferromagnet and the normal metal except at the interface. Writing the voltage drop at the interface (chosen as at $x=0$) by $V_{\rm FN}$, we obtain \begin{align} \nabla\cdot{\bm E} \simeq \delta(x) V_{\rm FN}/d, \end{align} where $d$ is the width of the interaface, which is treated as small enough compared with the electron mean free path, resulting in the delta function in $\nabla\cdot{\bm E}$. In totally unpolarized non-magnetic metals, namely, if $\sigma_+=\sigma_-$ and $D_+=D_-$, the correlation function in \Eqref{chidef} always vanishes. As is naively guessed, therefore, spin injection thus requires an effective spin polarization close to the interface, induced by the exchange interaction with the ferromagnet. This spin polarization is expected to be localized within a short distance of a few lattice constants from the interface. Let us approximate the interface polarization by introducing spin-dependent diffusion constant and the density of states, $ \overline{D}_\sigma$ and $\overline\nu_\sigma$, respectively, at the interface. The long-range behavior of the spin correlation function in the non-magnetic side is then obtained as \begin{align} \chi^{\rm (N)}({\bm x}) &= -\frac{e}{4\pi} (\sum_\sigma \sigma \overline{D}_\sigma\overline\nu_\sigma)\frac{e^{-|{\bm x}|/\ls{}}}{|{\bm x}|}, \end{align} where $\ls{}$ in the spin diffusion length in the normal metal ( $\ls{}$ is a long ($\sim \mu$m) length scale and thus does not depend on the spin). We therefore obtain from \Eqref{muspin} the chemical potential as \begin{align} \mu_{\rm s}^{\rm (N)} ({\bm x}) &= \frac{q_{\rm s} }{4\pi |{\bm x}|}e^{-|{\bm x}|/\ls{}}, \end{align} where $q_{\rm s}\equiv (\sum_\sigma \sigma \overline{D}_\sigma \overline\nu_\sigma) eV_{\rm FN})A_{\rm FN}/d $, is the spin accumulation rate at the interface (per unit time), and $A_{\rm FN}$ is the area of the junction. This result of $\mu_{\rm s}$ is consistent with intuitive and phenomenological results of the spin injection in the perpendicular structure shown in Fig. \ref{FIGspininjection}(a). In contrast, when the voltage is applied parallel to the ideal interface as shown in Fig. \ref{FIGspininjection}(b), spin injection does not occur since $\nabla\cdot{\bm E}=0$ at the interaface and thus $ \mu_{\rm s}^{\rm (N)}=0$. \begin{figure}[tbh] \begin{center} \includegraphics[width=0.5\hsize]{spininjection.eps} \caption{ (a) Creation of diffusive spin current (spin injection) by applying the electric voltage perpendicular to the F-N interface. (b) When the voltage is applied parallel to an ideal interface, no spin current is induced since $\nabla\cdot{\bm E}=0$. \label{FIGspininjection} } \end{center} \end{figure} \section{Total torque and asymmetric $\beta$ term} The continuity equation (\ref{sconteq}), indicates that the spin polarization (magnetization) changes due to the spin relaxation. (Change of the magnetization magnitude is a feature of the itinerant magnetism.) By use of Eqs. (\ref{TauE})(\ref{mudiffusioneq})(\ref{jsexpression2}), we see that \Eqref{sconteq} results in \begin{align} \dot{s}^z &= \gamma_{\tau} \nabla\cdot{\bm E}+\delta \gamma_{\tau} \nabla_z E_z +\sum_{\pm}(\pm)\frac{\mu_{\pm}}{(\ls{,\pm})^2}, \label{dotsz} \end{align} where $\gamma_{\tau} \equiv \gamma-\delta \sigma_{{\rm s}}$ and $\delta \gamma_{\tau} \equiv \delta\gamma+\delta \sigma_{{\rm s}}$. General case with uniform magnetization along any unit vector ${\bm n}$ is given by (${{\bm s}}\equiv s{\bm n}$) \begin{align} \dot{{{\bm s}}} = {\bm n} \left( \gamma_{\tau} \nabla\cdot{\bm E} + \delta \gamma_{\tau} \nabla_\parallel E_\parallel +\sum_{\pm}(\pm)\frac{\mu_{\pm}}{(\ls{,\pm})^2} \right), \label{dotsnv} \end{align} where $E_\parallel\equiv {\bm n}\cdot{\bm E}$ and $\nabla_\parallel\equiv {\bm n}\cdot\nabla$. In addition to the change of the magnitude, \Eqref{dotsnv}, there is a torque, which is perpendicular to ${\bm n}$. Such torque arises when the magnetization is not homogeneous, and plays important roles in current-induced magnetization dynamics. We have carried out the calculation of the current-induced torque done in Ref. \cite{TE08} on the same footing as the derivation of \Eqref{TauE}. As a result, we found that the $\beta$ term becomes asymmetric as (see Appendix \ref{APP:torque} for details of the calculation) \begin{align} {\cal T}^{(\beta),\alpha}_{{\rm so}} & = -\frac{P}{es^2} [\beta {{\bm s}}\times (\bm{j}\cdot\nabla){{\bm s}} +\delta \beta {{\bm s}}\times (j_\parallel \nabla_\parallel){{\bm s}}]^\alpha, \end{align} where $j_\parallel\equiv {\bm n}\cdot\bm{j}$ is the current along the local magnetization and $\delta \beta/\beta=-1/5$ in the present model. The spin transfer torque due to the spin-orbit interaction is thus different from that due to the spin-flip scattering. The expression of the total torque allowing for the spatially varying current density and the magnetization is therefore obtained as \begin{align} \dot{{{\bm s}}} &= -\frac{P}{2e} (\nabla\cdot\bm{j}){\bm n} -\frac{P}{e}\left[ \beta {\bm n}\times (\bm{j}\cdot\nabla){\bm n} +\delta \beta {\bm n}\times (j_\parallel \nabla_\parallel){\bm n}) \right] \nonumber\\ & + {\bm n} \left( \gamma_{\tau} (\nabla\cdot{\bm E}) +\delta\gamma_{\tau} (\partial_\parallel E_\parallel) +\sum_{\pm} (\pm) (\ls{,\pm})^{-2}\mu_\pm \right). \label{torquesum} \end{align} This expression clearly demonstrates that the spin relaxation torque requires some inhomogeneity either of the applied current or the spin structure, in addition to the spin-orbit (or spin flip) interaction. Totally homogeneous system does not relax. The last term in Eq. (\ref{torquesum}) gives useful information for measuring the spin accumulation induced by the spin current. \section{Conclusion} We have carried out a microscopic calculation of the spin relaxation torque and the spin current induced in disordered ferromagnetic metals by the applied electric field. The spin-orbit interaction arising from the random impurities is included as a source of spin relaxation, and inhomogeneity of the applied electric field is taken into account. We found that the spin relaxation torque in the uniform magnetization case is written as a divergence of the electric field plus an anisotropic term. The spin current was shown to be made up of field-driven (local) and diffusive (nonlocal) contributions, the latter written as a gradient of a spin chemical potential. We have derived a general linear response expression for the spin chemical potential. The spin injection effect was briefly discussed based on our results. When the analysis is applied to the inhomogeneous magnetization case, we argued that the $\beta$ torque in the current-induced magnetization dynamics can be anisotropic. Before finishing, we emphasize that the expression for the spin current and $\mu_{\rm s}$ are meaningless without specifying the physical observable to be measured. In the inverse spin Hall effect, which was originally proposed as \cite{Hirsch99,Takahashi08,Saitoh06} $j_\mu \propto \epsilon_{\mu\nu\rho}\jspin{,\nu}{\rho}$, it has recently been demonstrated that the charge current is not directly proportional to the spin current \cite{Hosono10,Takeuchi10}. Solving for the spin current only does not therefore provide physical information. \section*{Acknowledgment} The authors thank E. Saitoh, J. Shibata, H. Kohno, S. Murakami for valuable discussions. This work was supported by a Grant-in-Aid for Scientific Research in Priority Areas, "Creation and control of spin current" (1948027), the Kurata Memorial Hitachi Science and Technology Foundation and the Sumitomo Foundation.
\section{Introduction} According to the hierarchical model of structure formation, clusters of galaxies are the most massive objects in the universe and the cluster mass function is a powerful probe of cosmological parameters (e.g. \citealt{evrard89,eke98,henry00,allen04,vikh09}). In addition, the ratio between the cluster gas mass, as estimated with X-ray observations, and the total mass in a galaxy cluster provides stringent constraints on the total matter density. Specifically, the apparent evolution of the gas fraction with redshift can be used to estimate the contribution of the dark energy component to the cosmic density (e.g \citealt{allen08,ettori09}). The use of clusters as a cosmological probe therefore requires reliable mass estimates. \\ Several techniques are commonly used to estimate masses for galaxy clusters: the X-ray luminosity or temperature of the hot intracluster gas, the Sunyaev-Zel'dovich effect, the number of bright galaxies in a cluster, and the velocity dispersion of the cluster galaxies. The disadvantage of all of these methods is that they are indirect and require significant assumptions about the dynamical state of the cluster. Gravitational lensing, in contrast, is only sensitive to the amount of mass along the line of sight and allows reconstruction of the projected cluster mass regardless its composition or dynamical behavior (e.g. \citealt{KS93}). The only direct method to estimate cluster masses is therefore via measurement of the distortion (\textit{shear}) of the shapes of background galaxies that are weakly lensed by the gravitational potential of the cluster.\\ This distortion is very small and lensed galaxies are usually at high redshift. Observational studies to measure weak gravitational lensing by clusters require deep images in order to detect these faint sources and to obtain a high number density of background galaxies. Moreover this kind of analysis requires very high quality images in order to measure the shape of the lensed sources with high precision: good seeing ($ <1\arcsec$) conditions and a high signal-to-noise ratio (typically $>10$) are needed. Wide-field images are also required to obtain a statistical measure of the tangential shear as function of distance from the cluster center so that the projected mass measured by weak lensing includes essentially all of the mass of the cluster (\citealt{clowe01,clowe02}).\\ In the last decade substantial progress has been made with weak lensing studies thanks to the advent of wide-field data with linear detectors, the development of sophisticated algorithms for shape measurements (e.g. \citealt{KSB95}, \citealt{BJ02}, \citealt{refregier03}, \citealt{kui06}), and the availability of multi-band photometry which provides information about the redshift distribution of the lensed sources \citep{Ilbert}. \\ Here we describe the results of a weak lensing analysis of the \object{Abell 611} cluster. This analysis is based on images obtained with the Large Binocular Camera (LBC), which are a pair of prime focus cameras mounted on the two 8.4m diameter mirrors of the Large Binocular Telescope (LBT). Each LBC has a $23'\times25'$ field of view (FOV) and, combined with the collecting area of LBT, is a very powerful instrument for weak lensing studies. \\ Abell 611 is a rich cluster at redshift $z=0.288$ \citep{red} that appears relaxed in X-ray data, has a regular morphology, and the brightest cluster galaxy (BCG) is coincident with the center of X-ray emission (Donnarumma et al. in preparation). A giant arc due to strong lensing is also clearly visible close to the BCG (Fig. \ref{fig:ugr}). \begin{figure} \begin{center} \includegraphics[width=8cm]{A611_ugr_or.ps} \caption{A three-color image ($2'\times1.6'$) of Abell 611 obtained with LBC observations in the $u$-, $g$-, and $r$-bands. A giant arc is clearly visible close the brightest cluster galaxy ($\alpha$= 08h 00m 57s,$\delta$=+36d 03' 23'').}\label{fig:ugr} \end{center} \end{figure} \\ In this work we describe a weak lensing analysis to estimate the mass of Abell 611 from a deep $g$-band LBC image whose field of view extends well beyond the expected virial radius of the cluster. We compare the mass estimated from gravitational lensing with previous lensing results and with other mass estimates available in the literature that were derived by secondary techniques. In particular we compared our weak lensing results with new mass measurements obtained by X-ray analysis of Chandra data provided to us by Donnarumma et al. (in preparation).\\ Mass measurements by weak lensing do not need any assumption about the geometry of the cluster; however, assumptions are required to compare projected lensing masses with other mass estimates. Thus projection effects have to be taken into account during this kind of analysis, in particular the true triaxiality of the halo (\citealt{defilippis05}, \citealt{gavazzi05}) and the presence of unrelated structures along the line of sight (\citealt{metzler}, \citealt{hoekstra07}) can be sources of noise or bias on the projected mass measurements.\\ The paper is organized as follows. In the first sections we describe the data (Section \ref{data}) used for a weak lensing analysis of Abell 611, the catalog extraction of the background sources (Section~\ref{catalog}), and the selection of candidate cluster galaxies (Section \ref{red}). The two different pipelines used to extract the shear signal from the images are described in Section \ref{analysis}, and their results are compared in Section \ref{comparison}. Finally, both shear maps are used to estimate the mass of the cluster with different techniques (Section \ref{mass}). The results are summarized and discussed in the Section \ref{summary}.\\ Throughout this paper we adopt $H_0$= 70 km $s^{-1}$ $Mpc^{-1}$, $\Omega_{m}$=0.3, and $\Omega_{\Lambda}$=0.7. At the distance of Abell 611, $11'$ radius corresponds to a projected physical distance of nearly 3 Mpc. \section{Observations and Data Reduction}\label{data} Abell 611 was observed in March 2007, during the Science Demonstration Time (SDT) for the blue-optimized Large Binocular Camera (LBC), which is one of the two LBCs built for the prime foci of the LBT. The LBC focal plane consists of four CCDs (2048 x 4608 pixels, pixel scale 13.5 $\mu$m, gain $\sim$ 2 e$^-$/ADU, readout-noise $\sim11$ e$^-$). The CCDs are arranged so that three of the chips are butted along their long edges and the fourth chip is rotated counterclockwise by 90 degrees and centered along the top of the other CCDs (see Fig.\ref{fig:box}). The field of view is equivalent to $23'\times25'$ and provides images with a sampling of $0.225''$/pixel. Because each LBC is mounted on a swing arm over the primary mirror, the support structure lacks the symmetry of most prime focus instruments. Moreover LBC PSFs are dominated by optical aberrations from mis-alignments, which can cause PSFs to not have bilateral symmetry. This is a potential complication for the weak lensing analysis and we discuss this point further below. More details about the characteristics of LBC are given in \citet{giallongo}.\\ The observations, collected in optimal seeing conditions (FWHM $\sim 0.6''$), consisted of several sets of exposures of 5 minutes each in a wide $u$-band, SDSS $g$- and $r$-band filters. The total exposure time was 1 hour in $g$, 15 minutes in $r$, and 30 minutes in $u$. Unfortunally some $u$-band observations were not usable, so the real total exposure time used for this band is 20 minutes. For the present work we used the deep, $g$-band data for the weak-lensing analysis and the $u$- and $r$-band data to select cluster galaxies. Each image was dithered by 5 arcseconds in order to remove bad pixels, rows, columns, and satellite tracks. This offset is not large enough to fill the gaps between the CCDs, but the analysis plan was to treat CCDs separately due to expected PSF discontinuities at the chip boundaries. The offsets were therefore kept small to maximize depth and uniformity. \begin{figure} \begin{center} \includegraphics[width=8cm]{A611_boxn.ps} \caption{$g$-band image of the full field of LBC, centered on Abell 611. The box marks the region of 5000 pixels we used for the analysis.}\label{fig:box} \end{center} \end{figure} The images were reduced by the LBT pipeline\footnote{http://lbc.oa-roma.inaf.it} implemented at INAF-OAR. The flat-field correction was done using both a twilight flat-field and a superflat obtained during the night. Moreover a geometric distortion correction was performed in order to normalize the pixel size, which showed differences across the CCDs due to field distortions in the optics. The astrometric solution was computed using the \textsc{ASTROMC} package \citep{rad08}. This solution was then used to resample and coadd the images using the \textsc{SWarp}\footnote{developed by E. Bertin, http://terapix.iap.fr} software. \\ Standard fields for photometric calibration were not observed during the SDT. So we used the values of zero points for each band (Tab. \ref{tab:filters}) given by \citet{giallongo}. Table \ref{tab:filters} also shows the limiting magnitudes estimated from the faintest point-like objects detected at the 5$\sigma$ and 20$\sigma$ level. \begin{table} \begin{center} \begin{tabular}{cccccc} \hline Filter & $N_{exp}$ & exptime & mag & mag & zero point \\ & & (s) & ($5\sigma$) & ($20\sigma$) & \\ \hline \hline g & 12 & 3600 & 28.0 & 26.0 & 28.4 \\ r & 3 & 900 & 26.0 & 24.2 & 27.7 \\ u & 4 & 1200 & 27.2 & 25.0 & 27.1 \\ \hline \end{tabular} \end{center} \caption{Exposure times, limiting magnitudes for point-like sources and zero points (AB) for the observations in each band. \label{tab:filters} } \end{table} \section{Catalog extraction}\label{catalog} The detection of sources was performed using the \textsc{SExtractor} package \citep{bertin96}. Regions of the image presenting potential problems, such as spikes and halos around bright stars, were masked by visual inspection; sources inside such regions were discarded from the final catalog. In addition, we removed sources located at the borders of each CCD, where the SNR was lower due to the small dither offset. Finally, a very bright star dominates one of the CCDs. We therefore decided to limit our analysis to a box (displayed in Fig.\ref{fig:box}) with a size of 5000 pixels (corresponding to $\sim18.7'$) centered on Abell 611. Starting from this box of 350 arcmin$^2$, the effective area used for the analysis was $\sim$ 290 arcmin$^2$ after removing all the masked regions (30\% due to bright stars, 70\% due to regions between adjacent CCDs with no data or low S/N).\\ \begin{figure} \begin{center} \includegraphics[height=8cm,angle=-90]{pcats.ps} \caption{Magnitude ($g$) vs. half-light radius ($r_h$) plane. \ \textit{Green zone} shows the background galaxies selected for the lensing analysis ($23<g<26$, $r_h>1.8$ pixels) and \textit{pink zone} the unsaturated stars selected for correction of PSF anisotropy ($20<g<23$, $1.4<r_h<1.8$ pixels). See \S\ref{catalog} for further details.}\label{fig.mag_rh} \end{center} \end{figure} The separation between stars and galaxies was performed in the $mag-r_h$ plane, where magnitudes (mag) and half-light radii ($r_h$) were obtained from the MAG\_AUTO and FLUX\_RADIUS parameters computed by \textsc{SExtractor}. Unsaturated stars were selected on the vertical branch (see Fig. \ref{fig.mag_rh}) in the range $20 < g < 23$ mag and $1.4<r_h<1.8$ pixels. In this way we obtained 302 stars for the PSF correction, with a SNR$>200$ for the faintest ones. \\ For the lensing analysis, only background galaxies located at redshifts larger than $z=0.288$, the redshift of the cluster, should be used. Unfortunately, in our case the number of available bands does not allow us to estimate accurate photometric redshifts of these faint galaxies. The selection of the background galaxies was therefore done by choosing an adequate cut in apparent magnitude.\\ The choice of the upper magnitude limit was based on the galaxy redshift distribution obtained by \citet{Ilbert} from the Canada-France-Hawaii Telescope Legacy Survey (CFHTLS), which also used the SDSS photometric system. Taking into account the accuracy of the photometric redshifts (3\%) of \cite{Ilbert}, the approximations due to the different bands they used compared to ours and the assumption that we have the same galaxy distribution in our field, we chose $0.4$ as redshift reference value to perform the magnitude cut. Galaxies with $z \leq 0.4$ were assumed to belong to the cluster or be foreground galaxies. This reference value was chosen to be larger than cluster redshift, in order to reduce the contamination of foreground galaxies as much as possible, taking into account the approximations discussed above. Figure \ref{fig_CFHTLS} shows the fraction of the total CFHTLS sources at $z \leq 0.4$ ({\it upper panel}) and the fraction of background galaxies at $z > 0.4$ ({\it bottom panel}) as a function of the apparent magnitude cut. From this figure we conclude that a magnitude cut at $g>23$ is a good compromise to minimize the contamination from likely foreground and cluster galaxies ($\sim10$\%) and to maximize the number density of background galaxies ($\sim98$\%). The faint magnitude cut was chosen at $g<26$, that is the magnitude limit where we have a signal-to-noise ratio SNR $>10$ for the sources, where SNR is defined as FLUX/FLUX\_ERR as measured by \textsc{SExtractor}. The final catalog contained 8134 background galaxies. \begin{figure} \begin{center} \includegraphics[width=8cm]{Isto_N_CFHTLS_tris.ps} \caption{The fraction of likely foreground and cluster galaxies (\textit{top-panel}) and background galaxies (\textit{bottom-panel}) from the CFHTLS versus magnitude cut in $g$-band. The dotted vertical line is the lower limit of the magnitude cut adopted to select background galaxies. See \S\ref{catalog} for details.}\label{fig_CFHTLS} \end{center} \end{figure} \section{Candidate cluster members}\label{red} The candidate cluster members were selected from the simultaneous usage of $u$-, $g$- and $r$- band photometry. To this end catalogs were extracted from these images, running SExtractor in dual-mode with the $g$-band image as detection image. Only sources detected in all three bands were used. We then applied an algorithm (Fu et al., in preparation) similar to the C4 Clustering Algorithm \citep{Miller05}. The algorithm is based on the assumption that galaxies in a cluster should have similar colors and locate together in space. It evaluates the probability of each galaxy to be field-like: hence candidate cluster galaxies are those for which this probability is below a given threshold, as outlined below. \begin {enumerate} \item Each galaxy was set in a four-dimensional space of $\alpha$, $\delta$, $u-g$ and $g-r$. For each galaxy (named ``target galaxy''), we counted the number of neighbors within the four-dimensional box, $N_{\rm target}$. The angular size of the box was set to 1 $h^{-1}$ Mpc ($\sim 5.5'$ at the redshift of Abell 611). The sizes of the boxes in two color dimensions were determined as \begin{equation} \delta_{mn} = \sqrt{\sigma_{mn}^2({\rm stat}) + \sigma_{mn}^2({\rm sys})}, \end{equation} where $\sigma_{mn}({\rm stat})$ is the observed error for two magnitudes $(m,n)$, and $\sigma_{mn}({\rm sys})$ is the intrinsic scatter of the color $m-n$. For LBC A611 data, $ \delta_{ug}$, $ \delta_{gr}$ are 0.49 and 0.31 respectively. \item This four-dimensional box was placed on 100 randomly chosen galaxy positions and at each position we counted the number of neighbors. These randomized number counts constructed a distribution of counts for the target four-dimensional box. This distribution is represented by the median value $N_{\rm median}$ of the randomization counts. \item The probability $p$ of the target galaxy being field-like was derived by comparing the target galaxy count $N_{\rm target}$ to the distribution of randomization values $N_{\rm median}$. \item The distribution of $p$ values was derived by repeating the above steps for all galaxies. We ranked the $p$ values from smallest to highest and derived the value after which $p$ starts to rise significantly. In this way we identified $\sim$ 150 galaxies at $r < 23$ mag as the candidate cluster members. We further removed outliers in the $g-r$ vs $u-g$ diagram, leaving 125 candidate members. \end {enumerate} Starting from these galaxies, we fitted the $g-r$ vs. $r$ Red Sequence ($g-r = a+b\cdot r$) using a biweight regression method (Fig. \ref{fig:red_seq}), and obtained: $a= 2.39_{-0.45}^{+0.61}$, $b=-0.04_{-0.03}^{+0.02}$. \begin{figure} \centering \includegraphics[width=7cm,angle=-90]{redseq.ps} \caption{Colour-magnitude plot of the galaxies in the Abell 611 field. Red points are the candidate cluster galaxies selected by the C4 method. Also displayed (solid line) is the results of the biweight regression fit.}\label{fig:red_seq} \end{figure} \section{Weak lensing analysis}\label{analysis} Weak lensing is based on the measurement of the coherent distortion of the shapes of background galaxies produced by a distribution of matter. This distortion is very small and it is unmeasurable for any single background galaxy because the galaxy intrinsic ellipticity is not known. The dispersion of intrinsic ellipticity is a source of noise which is $\propto \sigma_\epsilon/\sqrt{N}$. A statistical approach is therefore required, where the distortion can be measured for a large number of sources in order to bring down that noise. This requires a careful treatment of systematic effects, as the shapes of galaxies may also be affected by contributions to the point spread function (PSF) by both the telescope and the atmosphere. In the last decade several methods have been developed for this kind of analysis. The most popular is the KSB approach, originally proposed by \citet{KSB95} and improved by \citet{LK97} and \citet{hoekstra98}. Several different implementations of this method exist in the literature and have been used in a number of KSB analysis pipelines. More recently, \citet{refregier03} and \citet{MR05} have proposed a new method based on Shapelets. Several available pipelines have also used this approach to measure the shear signal in various ways (e.g. \citealt{MR05}, \citealt{kui06}). \newline One of the key differences between these two approaches is the treatment of the PSF. The KSB method assumes that the PSF can be written as a convolution of a very compact anisotropic kernel with a more extended, circular function. These two terms are expressed in terms of the quadrupole moments of the surface brightness. As summarized below, the isotropic component is subtracted from the measured ellipticity and the anisotropic component is subtracted from a responsivity term. In contrast, in the Shapelets approach there is no assumption about the PSF shape. Individual galaxy images are decomposed into a complete orthonormal basis set consisting of Hermite (or Laguerre) Polynomials and the PSF correction is performed through deconvolution. \\ Since the PSF of LBC presents a significant deviation from symmetry, the analysis of Abell 611 presents a very good opportunity to compare the results produced by these two methods. For this comparison we started with the same initial catalogs of stars and galaxies for both pipelines. Specifically, we only considered sources with a \textsc{SExtractor} FLAG $<$ 4, which removes sources that are possibly blended. As the subsequent steps performed by each algorithm are different, the same galaxy may be rejected by one algorithm but not by the other. The result is that different output catalogs are produced by the KSB and Shapelets pipelines. To have a homogeneous comparison, we therefore finally selected only the sources common to both output catalogs. Sources with an unphysical ellipticity $|e|>1$ were also removed from the final catalogs. \subsection{KSB method}\label{ksb} We used the weak lensing pipeline described in \citet{rad08} to compute the quantities relevant to the lensing analysis. This pipeline implements the KSB approach using a modified version of the IMCAT\footnote{http://www.ifa.hawai.edu/$\sim$kaiser/imcat} tools that was provided to us by T. Erben (\citealt{erben01} and \citealt{hetter07}). \newline In the KSB approach stars and galaxies are parametrized according to the weighted quadrupole moments of the intensity distribution using a Gaussian weight function whose scale length is the size of the source (the formalism is described in \citealt{KSB95}). The main assumption of this approach is that the PSF can be described as the sum of a large isotropic component (seeing) and small anisotropic part. In this way the observed ellipticity $e_{obs}$ can be related to the intrinsic source ellipticity $e_{s}$ and shear $\gamma$ by the relation: \begin{equation} e_{obs}=e_s+P^{\gamma}\gamma+P^{sm}p, \end{equation} where $P^{sm}$ is the smear polarizability tensor and $P^{\gamma}$ is the pre-seeing polarizability tensor, which is related to the shear polarizability tensor $P^{sh}$ and $P^{sm}$ by Eq.~\ref{pgamma} \citep[see also][]{hoekstra98}. The quantity $p$ characterizes the anisotropy of the PSF and is estimated from stars, which have zero intrinsic ellipticity: \begin{equation}\label{p} \centering p^{*}=\frac{e_{obs}^{*}}{P^{sm*}}. \end{equation} If we average over a large number of sources, assuming a random orientation of the unlensed galaxies, we expect $\langle e_s \rangle$ = 0 and so \begin{equation} \gamma \ = \ \langle\frac{e_{iso}}{P^{\gamma}}\rangle, \end{equation} where $e_{iso}=e_{obs}-P^{sm}p^{*}$ is the ellipticity corrected for anisotropic distortions.\\ \begin{figure*} \begin{center} \includegraphics[angle=-90,width=0.9\textwidth]{plpsfn.ps} \caption{Removal of PSF anisotropy in the KSB approach. The observed (\textit{top-left}), fitted (\textit{top-right}) and residual (\textit{bottom-left}) ellipticities of stars for all CCDs. The observed (black points) versus corrected (green points) ellipticities are also shown (\textit{bottom-right}).}\label{fig_psf} \end{center} \end{figure*} We computed the contribution due to the PSF anisotropy (Eq. \ref{p}) from the selected stars. This quantity changes with position in the image, so we needed to fit it on each CCD in order to extrapolate its value at the position of the galaxy we want to correct. In our case we performed this fitting on each CCD and a second-order polynomial fit was sufficient.\\ Fig. \ref{fig_psf} displays the spatial pattern of the ellipticities for the stars in all CCDs, before and after the PSF correction. \\ After that we computed $P^{\gamma}$ for each source: \begin{equation}\label{pgamma} P^{\gamma}= P^{sh}-P^{sm}\frac{P^{sh*}}{P^{sm*}}. \end{equation} As the PSF correction was done separately on each CCD, we decided not to fit $\frac{P^{sh*}}{P^{sm*}}$ as a function of the coordinates and instead we took an average value. \begin{figure} \begin{center} \includegraphics[width=7cm,angle=270]{pshpsm.ps} \caption{${P^{sh*}}/{P^{sm*}}$ values computed in different bins of $r_h$, for each CCD. These values are lower for the central CCD (\#2) where the quality of PSF is better.}\label{fig:pshsm} \end{center} \end{figure} Stellar ellipticities should be computed using the same weight function used for galaxies (see \citealt{hoekstra98}), so we considered a sequence of bins in $r_{h}$ and for each galaxy we selected the PSF correction terms computed in the closest $r_{h}$ bin. Fig. \ref{fig:pshsm} shows the $\frac{P^{sh*}}{P^{sm*}}$ values computed in different bins of $r_h$, for each CCD. \\ We weighted the shear contribution from each galaxy according to: \begin{equation}\label{w} w=\frac{P^\gamma{}^2}{P^\gamma{}^2 \sigma_{e0}{}^2 + \langle \Delta e^{2} \rangle}, \end{equation} as in \citet{hoekstra00}, where $\sigma_{e0} \sim 0.3$ is the intrinsic rms of galaxy ellipticities, $\langle\Delta e^{2}\rangle^{1/2}$ is the uncertainty in the measured ellipticity, where the average is computed on both components of ellipticity for each galaxy (see eq. A8, A9 in \citet{hoekstra00}). \\ A crucial point in this kind of study is the selection of the galaxies to use for the shear analysis, as the contamination of foreground galaxies can dilute the lensing signal and lead to an underestimate of the mass.\\ Since the PSF degrades somewhat at the borders of the image, we limited our analysis to an $18.7' \times 18.7'$ box centered around the BCG, as discussed in Section \ref{catalog}. After that we filtered the source catalog using the following criteria: $P^{\gamma}>0.25$, SNR $>10$, $r_h >$ 1.8 pixels, $23 < g < 26$, ellipticities smaller than one, obtaining a surface density of $\sim$ 25 galaxies/arcmin$^2$. \\ The cut $P^{\gamma}>0.25$ allowed us to discard sources that appeared too circular (e.g. stars incorrectly classified as galaxies). We considered only galaxies with SNR $> 10$ in order to avoid noisy objects, which can be a source of error in the computation of the shear signal. Finally, we used the magnitude cut $23 < g < 26$ to select background galaxies, whose choice was previously explained in Section \ref{catalog}. \subsection{Shapelets}\label{KK} Another approach for weak lensing analysis is the use of Shapelets, which are basis functions constructed from two-dimensional Hermite polynomials weighted by a Gaussian. The translation, magnification, rotation and shear of astronomical images can be expressed as matrices acting on Shapelets coefficients. The advantage of Shapelets is that a galaxy image can be described in reverse order: pixelation, convolution with the PSF, and finally distortion by shear. Shapelets have a free scale radius $\beta$ which is the size of the Gaussian core of the functions. Its truncated expansion describes deviations from a Gaussian over a particular range of spatial scales, which widens with order $N$. \\ For the analysis of the LBC data of Abell 611 we used the Shapelets pipeline developed by \citet{kui06}, starting with the same stars and galaxies used in the KSB method. Each star was first fit with a circular Gaussian and the median radius was computed. This radius was then multiplied by a factor of 1.3 for the Shapelets fits, which was found by \citet{kui06} to work well for a range of model PSFs up to Shapelets order $N=8$. Then we obtained a Shapelets description of the PSF for each star. In order to estimate the PSF model at the position of each galaxy, the Shapelets coefficients were interpolated by a fourth-order polynomial on the whole image frame (Fig. \ref{fig:psf_shape}). The residual of the PSF model fitting is shown in Fig.~\ref{fig:residual}. \\ The ellipticity of each source is then determined by least-squares fitting a model, which is expressed as the shear applied to a circular source to fit the object optimally. The extension order of Shapelets for galaxies is taken the same as for stars. Performing the least-squares fit, the minimum of $\chi^2$ can be found in a few Levenberg-Marquardt iterations \citep{Press86}. The errors of Shapelets coefficients for each source are derived from the photon noise, and these can be propagated through in the $\chi^2$ function. Thus the error of shear measurement $\sigma_\gamma$ is calculated from the covariance matrix which is given by the second partial derivatives of $\chi^2$ at the best fit. \\ The shear contribution from each galaxy is weighted according to: \begin{equation} w=\frac{ \sigma_{e0}^2}{\sigma_{e0}^2+\sigma_\gamma^2}, \end{equation} which combines the error in the shear measurements $\sigma_\gamma$ and the intrinsic scatter $\sigma_{e0}$. The last quantity was computed from the distribution of ellipticity components of galaxies.\\ Finally, we removed galaxies with SNR $<10$ and that failed the Shapelets expansion and radial profile cuts defined by \citet{kui06}. This eliminated galaxies that were not well detected or measured, providing a number density of $\sim 26$ galaxies/arcmin$^2$. Further details about the selection criteria are available in \citet{kui06}. \begin{figure*} \begin{center} \includegraphics[height=0.7\textwidth, angle=0]{psfmap.eps} \caption{ Shapelets PSF models interpolated at different positions on each CCD. The distribution of these models corresponds to their actual placement on the CCD mosaic, CCD1 to CCD3 from left to right in the bottom panel and CCD4 in the top. Contours are the representations of PSF shape decomposed by Shapelets. The X and Y values correspond to the pixel position.}\label{fig:psf_shape} \end{center} \end{figure*} \begin{figure*} \begin{center} \includegraphics[height=0.6\textwidth, angle=270]{psfstars_id42081_ccd1.ps} \includegraphics[height=0.6\textwidth, angle=270]{psfstars_id44549_ccd2.ps} \includegraphics[height=0.6\textwidth, angle=270]{psfstars_id34538_ccd3.ps} \includegraphics[height=0.6\textwidth, angle=270]{psfstars_id88419_ccd4.ps} \caption{ The star at the position near the center of each CCD are taken as an example to show the residual of PSF model fitting (left column). The real PSF shape and fitted PSF model are shown in the middle and right columns. }\label{fig:residual} \end{center} \end{figure*} \subsection{Comparison between KSB and Shapelets ellipticities}\label{comparison} After matching the two output catalogs, our final background galaxy sample has a surface density of $\sim$ 23 galaxies/arcmin$^2$.\\ In weak lensing studies, the background number density of ground-based telescope usually is around 15 to 20 galaxies/arcmin$^2$ (e.g. \citealt{PH07}). It depends not only on the size of telescope, the exposure time, the color filter, seeing condition, but also on galaxy selection criteria.\\ The maps of the PSF correction computed by both methods show a good-quality PSF in the central regions that then degrades further from the center of the field. Nevertheless the final correction is good with fluctuations in the PSF anisotropy less than 1\%. \begin{figure} \begin{center} \includegraphics[width=7cm]{ecom_new.eps} \caption{Comparison between the first component of the shear from the Shapelets and KSB methods (results with the second components are very similar).}\label{fig:g} \end{center} \end{figure} Figure \ref{fig:g} compares the first component of the shear $\gamma_1$ measured by the KSB and Shapelets methods. It shows good agreement in the range of $-0.5 < \gamma_1 < 0.5$. Some scatter is present for very elongated galaxies, but the fraction of these galaxies is less than 5\% in the final, common catalog. As shown in Fig.~\ref{fig:weight}, these strongly elongated galaxies are down-weighted nearly by a factor of 2 compared to small ellipticity galaxies, so that they do not affect the final mass measurements of the cluster (as described further below). A similar behavior is seen for the second component $\gamma_2$. These matched catalogs were used to compute estimates of the cluster mass as described in the next sections. \begin{figure} \begin{center} \includegraphics[width=10cm]{weight.KSB.nor.ps} \caption{The plot of the normalized galaxy weight as function of the absolute first component of ellipticity for KSB measurement. The applied normalization was $\Sigma w_i \times \delta e_1=1$, where $\delta e_1$ is the constant bin width of $e_1$. The Shapelets measurement shows similar down-weight for elongated ellipticity galaxies.}\label{fig:weight} \end{center} \end{figure} \section{Mass Measurements}\label{mass} The relationship between the shear $\gamma$ and the surface mass density is: \begin{equation} \gamma(\theta)=\frac{1}{\pi}\int_{R^2}d^2\theta^{\prime}D(\theta-\theta^{\prime})\kappa(\theta^{\prime}), \end{equation} where: $D=\frac{-1}{(\theta_1-i\theta_2)^2}$, $\kappa\equiv\Sigma/\Sigma_{cr}$ is the convergence, $\Sigma_{cr}=\frac{c^2}{4\pi G}\frac{D_s}{D_lD_{ls}}=\frac{c^2}{4\pi G}\frac{1}{D_l\beta}$ is the critical density of the cluster, and $D_s$, $D_l$ and $D_{ls}$ are source-observer, lens-observer and lens-source distances respectively. While weak lensing actually measures the reduced shear $g=\frac{\gamma}{1-\kappa}$, for $\kappa\ll1$, $g\sim\gamma$.\\ An optimal approach for computing the mass requires knowledge of the redshift of each background galaxy. As we do not have this information, we assumed that the background sources all lie at the same redshift according to the \textit{single sheet approximation} \citep{KS01}. An estimate of the redshift value to use for the weak lensing analysis was computed using the first release of photometric redshifts available for the D1 deep field of the CFHTLS, adopting the magnitude cut in the $g$-band chosen here for the background galaxies selection ($23 < g < 26$ mag; see Section~\ref{catalog}) and assuming a Gamma probability distribution \citep{gavazzi04}. This yielded a median redshift $z=1.05$ and for our analysis we assume all of the background galaxies lie at $z\sim1$, corresponding to $\langle\beta\rangle\ = \langle D_{ls}/D_s \rangle\ \sim \ 0.65$.\\ As discussed by \citet{hoekstra07}, the single-sheet approximation results in an overestimate of the shear by a factor \begin{equation} 1+\left[\frac{\langle\beta^2\rangle}{\langle\beta\rangle^2}-1\right] \kappa. \end{equation} We computed this factor using the CFHTLS catalog of photometric redshifts and obtained $\langle\beta^2\rangle/\langle\beta\rangle^2 = 1.167$.\\ \\ The convergence $\kappa$ gives an estimate of the surface mass density apart from an unknown additive constant -- the so-called \textit{mass-sheet degeneracy}. We tried to solve this degeneracy using two different approaches: assuming either that $\kappa$ vanishes at the borders of the image or a particular mass profile whose expected shear profile is known. \subsection{S-Maps}\label{s-map} In Figure \ref{fig:snr_ksb} we plot the so-called \textit{S-maps} \citep{schirm04} for these data. \textit{S-maps} are computed as the ratio $S=M_{ap}/\sigma_{Map}$ where: \begin{eqnarray} M_{ap}&=&\frac{\Sigma_ie_{t,i}w_iQ(|\theta_i-\theta_0|)}{\Sigma_iw_i}\\ \sigma_{M_{ap}}^2&=&\frac{\Sigma_ie_{t,i}^2w^2_iQ^2(|\theta_i-\theta_0|)}{ 2(\Sigma_iw_i)^2}, \end{eqnarray} For this calculation the image is considered as a grid of points, $e_{t,i}$ are the tangential components of the ellipticities of the lensed galaxies, which are computed by considering the center of each point of the grid, $w_i$ is the weight as defined in Equation \ref{w}, and $Q$ is a window function, chosen to be a Gaussian function defined by: \begin{equation} Q(|\theta-\theta_0|)= \dfrac{1}{\pi \theta_c^2}exp\left( -\dfrac{(\theta-\theta_0)^2}{\theta_c^2}\right) \end{equation} where $\theta_0$ and $\theta_c$ are the center and the size of the aperture. The ratio $S=M_{ap}/\sigma_{M_{ap}}$ provides an estimate of the SNR ratio of the dark matter halo detection. \textit{S-maps} are discussed further by \citet{schirm04}. We computed these maps using shear catalogs obtained from both the KSB and Shapelets pipelines. In both cases (see Fig. \ref{fig:snr_ksb}) the maps show that the lensing signal is peaked around the BCG, confirming that this is indeed the center of the mass distribution. The mass distribution also appears quite regular, which is in agreement with what is indicated by the X-ray maps. In Figure \ref{fig:red_map} S-map contours are overlaid on the $r$-band luminosity-weighted density distribution of the red sequence galaxies of Abell 611, selected in Section \ref{red}, showing that the mass distribution follows that of the red cluster galaxies. \begin{figure*} \begin{center} \includegraphics[viewport=60 20 470 400,width=7cm,clip]{snr_KSBn.ps}\includegraphics[viewport=60 20 470 400,width=7cm,clip]{snr_KKn.ps}\\ \caption{S-maps obtained from the KSB (\textit{left panel}) and the Shapelets (\textit{right panel}) analysis. The levels are plotted between $\sigma_{\rm min}=3.5$ and $\sigma_{\rm max}=5$. They are overplotted on a $g$-band greyscale image ($\sim 4$ arcmin) of the center of the field of Abell 611. }\label{fig:snr_ksb} \end{center} \end{figure*} \begin{figure} \centering \includegraphics[width=7cm,angle=-90]{ldist.ps} \caption{$r$-band luminosity-weighted density distribution of red sequence galaxies of Abell 611. The overlaid contour (black lines) is the S-map (computed and discussed in Section \ref{s-map}) showing the SNR of the shear signal around the cluster obtained from the Shapelets analysis. The levels are plotted between $\sigma_{\rm min}=3.5$ and $\sigma_{\rm max}=5$.}\label{fig:red_map} \end{figure} \subsection{Aperture densitometry} In order to trace the mass profile of the cluster, we computed the $\zeta$-statistic described in \citet{clowe98} and \citet{fahlman94}: \begin{eqnarray} \zeta(\theta_1)=\bar{\kappa}(\theta\le\theta_1)&-&\bar{\kappa}(\theta_2<\theta\le\theta_{max})=2\int^{\theta_2}_{\theta_1}\langle\gamma_T\rangle d\ln{\theta}\\ \nonumber &+&\frac{2}{1-(\theta_2/\theta_{max})^2}\int^{\theta_{max}}_{\theta_2}\langle\gamma_T\rangle d\ln{\theta}, \end{eqnarray}\label{apdens} Mass measurements are computed within different apertures of increasing radius using a control annulus far from the center of the distribution (the BCG). $M_{ap}=\pi\theta_1^2\zeta(\theta_1)\Sigma_{cr}$ is the mass within the last aperture and provides a lower limit on the mass, unless the value in the control annulus is equal to zero. This method allows us to choose the size of the annulus that satisfies the desired condition; moreover it has the advantage that this mass computation does not depend on the mass profile of the cluster \citep{clowe98}. \\ We chose $30''\leq \theta \leq 500''$ and $\theta_{max}=600''$, which yielded projected masses within $\sim 1500$ kpc of $7.7\pm{3.3}\times 10^{14} M_{\odot}$ and $8.4\pm{3.8}\times 10^{14} M_{\odot}$ using the KSB and Shapelets shear catalogs, respectively.\\ Unfortunately we could not extend our analysis further from the center of the cluster because of the presence of a very bright star in the field that made the outer regions unusable.\\ Since weak lensing is sensitive to the total mass along the line of sight, the observed aperture mass is the sum of the mass of the cluster and any contribution from other, uncorrelated structures along the line of sight. This contribution is assumed to be negligible in the central regions of the cluster, which are much denser, and become more relevant in its outer regions. As discussed by \cite{hoekstra01}, the effect of this contribution does not introduce any bias, but it does add a source of noise to the lensing mass. Aperture densitometry is more affected by this uncertainty than parametric methods because it is sensitive to the lensing signal at large radii. Nevertheless, for observations of rich clusters at intermediate redshifts, this uncertainty is fairly small because the bulk of the background sources are at much higher redshifts than the cluster. \subsection{Model fitting}\label{fit} The model fitting approach for estimating the mass of a cluster consists of assuming a particular analytic mass density profile for which calculate the expected shear and then fitting the observed shear with the model by minimizing the log-likelihood function (\citealt{schneider00}): \begin{equation} l_{\gamma}=\sum_{i=1}^{N_{\gamma}}\left[\frac{|\epsilon_i-g(\theta_i)|^2}{ \sigma^2[g(\theta_i)]}+2\ln\sigma[g(\theta_i)]\right],\label{log} \end{equation} with $\sigma[g(\theta_i)]=(1-g(\theta_i)^2)\sigma_e$.\\ In this analysis we assumed both a Singular Isothermal Sphere (SIS) and a Navarro-Frenk-White (\citealt{nfw}) model.\\ In the SIS model the density profile depends on one parameter, the velocity dispersion $\sigma$: \begin{equation} \rho_{SIS}(r)=\dfrac{\sigma^2}{2\pi G}\dfrac{1}{r^2}. \end{equation} In this profile the shear is found to be related to $\sigma$ by: \begin{equation} \gamma_T(\theta)=\frac{2\pi}{\theta}\frac{\sigma^2}{c^2}\frac{D_{ls}}{D_s}=\frac{\theta_E}{\theta}. \end{equation} (e.g. \citealt{BS01}).\\ The mass density profile predicted by the Navarro-Frenk-White model (hereafter NFW) is: \begin{equation} \rho_{NFW}(r)=\frac{\delta_c\rho_c}{(r/r_s)(1+r/r_s)^2} \end{equation} where $\rho_c=3H^2(z)/(8\pi G)$ is the critical density of the universe at the cluster redshift, $r_s$ is a scale radius related to the virial radius by means of the concentration parameter $c_{vir}=r_{vir}/r_s$ and $\delta_c$ is a characteristic overdensity of the halo: \begin{equation} \delta_c=\frac{\Delta_{vir}}{3}\frac{c^3}{\ln(1+c)-c/(1+c)}. \end{equation} where $\Delta_{vir}$ is the virial overdensity, approximated by $\Delta_{vir}\sim (18\pi^2+82(\Omega_{M}(z)-1)-39(\Omega_{M}(z)-1)^2)/\Omega_{M}(z)$ using the spherical collapse model \citep{BN98} for flat cosmologies.\\ The mass of the halo is: \begin{equation} M_{vir}=\frac{4}{3}\pi\Delta_{vir}\rho_{m} r_{vir}^3. \end{equation} where $\rho_{m}= \rho_{c}\cdot \Omega_{M}(z)$ is the mean density at the cluster redshift. We solved for the mass of the cluster with the expression for the shear $\gamma_{T}(r)$ derived by \citet{bart96} and \citet{WB00} and minimized Equation \ref{log} with the \textsc{MINUIT} package.\\ Table~\ref{tab:results} shows the best-fit parameters and the mass values derived by model fitting. The NFW profile was used keeping both the concentration and mass as free parameters (marked as NFW), and by using the relation between $c_{\rm vir}$ and $M_{\rm vir}$ proposed by \citet{bullock01} (hereafter MNFW): $c_{vir}=\frac{K}{1+z}\left(\frac{M_{vir}}{M_*}\right)^{\alpha}$, where $M_* = 1.5\times 10^{13} /h\ M_{\odot}$, $K=9$ and $\alpha =-0.13$.\\ \begin{table*} \begin{center} \begin{tabular}{llll|lll} \hline & & KSB & & & Shapelets & \\ \hline Parameter & MNFW & NFW & SIS & MNFW & NFW & SIS \\ \hline \hline $M_{200} (10^{14})$& $5.3_{-1.2}^{+1.4}$ & $5.6_{-2.7}^{+4.7}$ & & $5.3_{-0.8}^{+0.8}$ & $5.9 _{-1.7}^{+2.2}$ &\\ $r_{200}$ (kpc) & $1513_{-123}^{+119}$ & $1545_{-306}^{+345}$& & $1516_{-78}^{+77}$ & $1570_{-170}^{+177}$ & \\ $c_{\rm vir}$ & 4.51 & $3.9_{-2.1}^{+5.6}$ & & 4.50 & $3.7_{-1.3}^{+2.2}$ & \\ $\sigma_{cl}$ (km/s)& & & $778_{-27}^{+26}$ & & & $781_{-27}^{+26}$ \\ \hline $red. \chi^2$ & 1.46 & 2.02 & 1.94 & 1.62 & 2.12 & 2.31 \\ $Q$ & 0.21 & 0.11 & 0.10 & 0.17 & 0.10 & 0.06 \\ \hline \end{tabular} \end{center} \caption{Mass values computed for $M_{200}$ by model fitting using a SIS profile, a NFW profile and a constrained NFW (MNFW), according with \cite{bullock01}. Best fit parameters and reduced $\chi^2$ are listed for each fit performed using both KSB and Shapelets shear catalogs, together with the goodness of fit probability $Q$. \label{tab:results} } \end{table*} These values were computed assuming spherical symmetry for the cluster halo. The effect of departures from spherical symmetry (e.g. triaxial halos) on the determination of the total cluster mass have been studied by several authors (e.g. \citealt{gavazzi05}, \citealt{defilippis05}). \citet{defilippis05} showed that these effects are negligible when the mass is computed at large distances from the cluster center, although they are important at small radii. The same authors tried to recover a three-dimensional reconstruction of Abell 611 through a combined analysis of X-ray and Sunyaev-Zel'dovich observations and concluded that the cluster was approximately spherical, supporting our symmetry assumption. \\ \begin{figure*} \begin{center} \includegraphics[width=7cm,angle=270]{mKSB.ps}\includegraphics[width=7cm,angle=270]{mKK.ps}\\ \caption{Results of fitting by a NFW model ({\it dashed lines}), a constrained (M)NFW ({\it solid lines}) model \citep{bullock01} and a SIS profile ({\it dotted lines}), with the KSB ({\it left panel}) and Shapelets ({\it right panel}) pipeline. Average values of tangential ({\it black}) and radial ({\it red}) components of the shear, computed in logarithmic scale bins, are also plotted. Cluster masses were computed at $r_{200}$, estimated to be $\sim 350$ arcsec.}\label{fig_fit} \end{center} \end{figure*} In Figure~\ref{fig_fit} the results of fitting by a NFW model ({\it dashed lines}), a constrained (M)NFW ({\it solid lines}) model \citep{bullock01} and a SIS profile ({\it dotted lines}) are displayed, for the KSB ({\it left panels}) and Shapelets ({\it right panels}) pipeline, respectively. The black lines represent the best fit to the unbinned data as function of distance from the center of the cluster. Average values of tangential ({\it black}) and radial ({\it red}) components of the shear, computed in logarithmic scale bins, are also plotted. The latter components are expected to be zero in the absence of systematics errors.\\ In Fig. \ref{ll_NFW} the confidence contours for the NFW profile are plotted in a plane $r_s$ vs $c$: the levels show confidence at 68\% and 90\%, starting with the innermost one.\\ \begin{figure*} \begin{center} \includegraphics[width=7cm,angle=270]{ll_KSB.ps}\includegraphics[width=7cm,angle=270]{ll_KK.ps}\\ \caption{The confidence contours for NFW profile are plotted in the ($r_s$,$c$) plane, for KSB (\textit{left panel}) and Shapelets (\textit{right panel}), respectively. The confidence levels are at 68\% and 90\%, starting with the innermost one. The triangle shows the best-fit position.}\label{ll_NFW} \end{center} \end{figure*} The goodness of each fit ($\chi^2$) and its probability ($Q$) are also listed in Table \ref{tab:results}. \section{Discussion}\label{summary} We have conducted a weak lensing analysis of Abell 611 with deep $g$-band images from the LBC. Due to the complexity of the LBC PSF, we decided to use both the KSB and Shapelets methods to measure galaxy shapes and extract the shear signal. KSB parametrizes the sources using their weighted quadrupole moments and is based on a simplified hypothesis of a nearly circular PSF. In contrast, Shapelets uses a decomposition of the images into Gaussian-weighted Hermite polynomials and does not make any assumption about the best PSF model. Due to the large collecting area of LBT and the wide field of the LBC, we were able to extract a high number density of $\sim 25$ background galaxies per arcmin$^{2}$ over a wide field of $19\times19$ arcmin$^2$ and this allowed us to perform an accurate weak lensing analysis.\\ The two shear catalogs, derived by the KSB and Shapelets pipelines, were matched and common sources (with a number density of $\sim$ 23 galaxies/arcmin$^2$) were used to estimate mass measurements for Abell 611 by two different weak lensing techniques: aperture densitometry and a parametric model fitting. In both approaches we assumed that the BCG was the center of the mass distribution, which is supported by S-Maps (see Fig. \ref{fig:snr_ksb}, Section~\ref{s-map}).\\ The projected mass values obtained by aperture densitometry within a radius of $\sim1500$ kpc are: $7.7\pm{3.3}\times 10^{14} M_{\odot}$ and $8.4\pm{3.8}\times 10^{14} M_{\odot}$, using KSB and Shapelets, respectively. These estimations are model independent \citep{clowe98}, but they are affected by large uncertainties. As discussed in Section \ref{apdens}, the contribution of uncorrelated structures along the line of sight can be source of noise for aperture measurements, although they can be decreased by averaging the results for several clusters (\citealt{hoekstra01}) or corrected by using photometric redshifts of the sources, if available. \\ Table~\ref{tab:results} shows the results of fitting the observed shear with a parametric model. We assumed both a SIS and a NFW mass density profile. We first fitted a NFW model leaving both $c$ and $r_s$ as free parameters: as displayed in Fig. \ref{ll_NFW}, the uncertainty on these two parameters provided by the fit is high and does not allow to put a strong constrain on the concentration. Nevertheless, the best-fit value ($c \sim 4$) is in good agreement with the one obtained when the Bullock et al. (2001) relation is adopted ($c = 4.5$). For the NFW fits the quoted masses are $M_{200}$, within the radius $r_{200}$ where the density is 200 times the critical density. The estimated value of $r_{200}$ for Abell 611 is $\sim1.5$ Mpc ($5.8'$).\\ The weak lensing mass measurements obtained from both the KSB and Shapelets shear catalogs are in agreement, within uncertainties: by using a (M)NFW profile we obtain $M_{200} \sim 5.3_{-1.2}^{+1.4} \times 10^{14} M_\odot$ and $5.3_{-0.8}^{+0.8} \times 10^{14} M_\odot$, respectively. The smaller uncertainties of Shapelets results show that this method provide a higher accuracy than KSB. Moreover, the goodness of the fits ($\chi^2$) in Table~\ref{tab:results} shows that both Shapelets and KSB provide a best-fit with higher probablity ($Q$) by using a NFW mass density profile than a SIS profile. \\ Abell 611 was previously targeted for a weak lensing study by \citet{dahle06}, who used $V$ and $I$ observations from several facilities to target a large number clusters (see \citealt{dahle02} for more details). They used the KSB \citep{KSB95} shear estimator as described in \citet{kaiser00} to derive the shear signal from the images. The mass of the cluster was then derived by fitting the observed shear with a NFW mass density profile \citep{nfw}. They assumed a concentration parameter as predicted by \citet{bullock01} and obtained a mass value of $M_{180} = (5.21\pm 3.47)\times10^{14}h^{-1}M_{\odot}$ within $r_{180}$, the radius within which the density is 180 times the critical density. A more recent mass estimate of Abell 611 is presented in \citet{PD07}, who used the data collected in \citet{dahle06} and obtained a value of $M_{500} = 3.83\pm 2.89\times10^{14}h^{-1}M_{\odot}$ within $r_{500}$, the radius within which the density is 500 times the critical density. In these papers the authors extrapolated the NFW profile out to $r_{500}$ because the data were insufficient to extend this far in projection from the cluster center. For their work on Abell 611 $r_{fit}/r_{500}=0.59$ (note $r_{500}=0.66\times r_{200}$). \\ Our weak lensing estimates for the mass of Abell 611 are in agreement with the previous results of \cite{dahle06} and \cite{PD07}, but the depth and the larger area covered by LBC data allowed us to perform a more accurate analysis.\\ A recent weak lensing analysis of Abell 611 has recently been done by \cite{okabe09} using Subaru/Suprime-Cam observations in two filters ({\it i'} and {\it V}). The authors used the color ($V-i'$) information to select the background galaxies to use for their cluster lensing analysis, getting a galaxy number density of $\sim 21$ galaxies/arcmin$^2$, with a SNR $\sim 10$. By fitting the mass density profile of the cluster using a NFW model, they found a mass value $M_{200}= 5.13^{+1.16}_{-1.00}\times 10^{14}M_\odot$ and a $c_{vir}=4.14^{+1.73}_{-1.21}$ (private communication), which are in good agreement with our results.\\ A new mass estimation of Abell 611 was performed by \cite{newman09} over wide range of cluster-centric distance (from $\sim$ 3 kpc to 3.25 Mpc) by combining weak, strong and kinematic analysis of the cluster, based on Subaru, HST and Keck data, respectively. They found a mass value $M_{200}= 6.2^{+0.7}_{-0.5}\times 10^{14}M_\odot$ with $c=6.95\pm{0.41}$ by using a NFW model fitting, in agreement with our results. We note that such a large value of $c$ cannot be rejected by our NFW fits, due to the large uncertainties on NFW parameters (see Tab.\ref{tab:results} and Fig.\ref{ll_NFW}). \\ Finally, we also compared the mass values obtained by our weak lensing analysis to X-ray estimates of $M_{200}$ available in literature. \cite{SA07} analyzed Chandra data of several clusters and modelled their total mass profile (dark plus luminous matter) using a NFW profile. They found for Abell 611 a scale radius $r_s = 0.32_{-0.20}^{+0.10}$ Mpc and a concentration parameter $c = 5.39_{-1.51}^{+1.60}$, which provide a mass value $M_{200} \sim 8\times 10^{14}M_\odot$ at $r_{200}\sim 1.7$ Mpc, in agreement, within the uncertainties, with the results obtained by us. A more recent X-ray analysis of Chandra observations of Abell 611 has been performed by Donnarumma et al. (in preparation). They obtained a value of $M_{200}= 1.11\pm 0.21\times 10^{15}M_\odot$ at $r_{200}\sim 1900$ kpc, with $r_s = 407_{-86}^{+120}$ kpc and $c = 4.76_{-0.78}^{+0.87}$. Their projected mass at $r_{200}$ is $1.28 \pm 0.24 \times 10^{15}M_\odot$. Such value is in agreement, within the statistical uncertainties, with the mass measured by aperture densitometry, but higher than the value estimated by the parametric model. Additional information on the mass will be derived from a strong lensing analysis of Abell 611 (Donnarumma et al. in preparation). \\ This work shows that LBT is a powerful instrument for weak lensing studies, but we want to stress that the data here analyzed did not allow us to use the full capabilities of the telescope. The presence of bright saturated stars hampered to use the whole field of the camera for the analysis of Abell 611. In addition, the analysis of weak lensing is expected to be improved by the usage of the Red Channel in LBC, which was not yet available during these observations. The present results are nevertheless important to demonstrate the capabilities of LBC for weak lensing; we therefore plan to extend to other clusters such analysis, now using the Red Channel.\\ \begin{acknowledgements} Observations were obtained with the Large Binocular Telescope at Mt. Graham, Arizona, under the Commissioning and Science Demonstration phase of the Blue Channel of the Large Binocular Camera. The LBT is an international collaboration among institutions in the United States, Italy and Germany. LBT Corporation partners are: The University of Arizona on behalf of the Arizona university system; Istituto Nazionale di Astrofisica, Italy; LBT Beteiligungsgesellschaft, Germany, representing the Max-Planck Society, the Astrophysical Institute Potsdam, and Heidelberg University; The Ohio State University, and The Research Corporation, on behalf of The University of Notre Dame, University of Minnesota and University of Virginia.\\ This paper makes use of photometric redshifts produced jointly by Terapix and the VVDS team. \\ Part of the data analysis in this paper was done using the R software (http://www.R-project.org).\\ We thank the anonymous referee for useful comments and suggestions which improved the presentation of this work. AR acknowledges the financial support from contract ASI-COFIS I/016/07/0. LF, KK and MR acknowledge the support of the European Commission Programme 6-th framework, Marie Curie Training and Research Network “DUEL”, contract number MRTN-CT-2006-036133. LF is partly supported by the Chinese National Science Foundation Nos. 10878003 \& 10778725, 973 Program No. 2007CB 815402, Shanghai Science Foundations and Leading Academic Discipline Project of Shanghai Normal University (DZL805). AD, SE, LM, MM acknowledge the financial contribution from contracts ASI-INAF I/023/05/0 and I/088/06/0. SPH is supported by the P2I program, contract number 102759.\\ \end{acknowledgements} \bibliographystyle{aa}
\section{Introduction} Recently geometric Langlands duality has been interpreted \cite{KW} as a consequence of an isomorphism between two different topological gauge theories in four dimensions whose gauge groups are related by Goddard-Nuyts-Olive duality \cite{GNO}. These topological gauge theories are obtained by twisting $N=4$ $d=4$ super-Yang-Mills theory by means of the so called GL-twist. The goal of this paper is to classify and study certain nonlocal observables (surface operators) in GL-twisted gauge theory, with a view towards strengthening and extending the geometric Langlands duality. From a mathematical point of view, a 4d Topological Field Theory (TFT) is a gadget which assigns a number to a compact oriented four-manifold, a vector space to a compact oriented three-manifold, a category to a Riemann surface, a 2-category to a circle, and a 3-category to a point. These data have varying amount of extra structure whose complexity is inversely related to the categorical level. For example, the category attached to a Riemann surface is acted upon by the mapping class group of the surface. The 2-category attached to a circle is a braided monoidal 2-category. The 3-category attached to a point is self-dual. From a physical viewpoint, the category attached to a Riemann surface $\Sigma$ can be interpreted as the category of boundary conditions for an effective 2d TFT obtained by compactifying the 4d TFT on $\Sigma$. The 2-category attached to a circle is the 2-category of boundary conditions for the 3d TFT obtained by compactifying the 4d TFT on the circle. Alternatively, it is the 2-category of surface operators in the 4d TFT. Equivalences of 4d TFTs therefore have implications for TFTs in lower dimensions. This was exploited in \cite{KW} where Montonen-Olive duality was shown to imply an equivalence of categories of boundary conditions for certain topological sigma-models in 2d. Going one dimension higher, we may consider GL-twisted $N=4$ gauge theories with gauge groups $G$ and ${{}^L G}$ compactified on a circle. The corresponding 3d TFTs are also isomorphic, so their 2-categories of boundary conditions must be equivalent. From the 4d point of view, the 2-category of boundary conditions in the 3d TFT is simply the 2-category of surface operators, where we ``forgot'' about the braided monoidal structure. It would be a mistake to think that the independence of the GL-twisted gauge theory on the metric means that we can let the circle radius to be zero and interpret the 3d TFT as a twist of $N=8$ $d=3$ super-Yang-Mills theory. This would imply that certain topological versions of $N=8$ $d=3$ SYM theories with gauge groups $G$ and ${{}^L G}$ are equivalent. However, this is not correct, as can be seen already in the abelian case. Indeed, compactifying $U(1)$ gauge theory on a circle of finite radius gives rise to a periodic scalar in the effective 3d theory (the holonomy of the gauge field along the compact direction). The period becomes very large as the circle radius goes to zero. But no matter how large the period is, the theory of a periodic scalar is different from that of an ${\mathbb R}$-valued scalar. For example, the former theory admits disorder loop operators where the scalar has a nonzero winding, while the latter theory does not. These disorder loop operators are crucial for maintaining the Montonen-Olive duality, as we will see below. In this paper we study the aspects of Montonen-Olive duality which can be understood in 3d terms. Namely, we study the 2-category of surface operators in the GL-twisted 4d theory, regarding it as a 2-category of boundary conditions in a 3d TFT. Some examples of surface operators in 4d TFT have been discussed in \cite{GW}, but we will see that there are much more general ones. The GL-twisted theory depends on a complex parameter $t$ which determines the BRST operator of the theory. For $t=i$ the analysis of surface operators is relatively simple, since the 3d theory turns out to be of a rather familiar kind (the gauged Rozansky-Witten model). In this case we propose a description of the full 2-category of surface operators. For $t=1$ and $t=0$ the 3d theory is rather unusual, and in this paper we analyze it only in the abelian case. We also describe how abelian electric-magnetic duality acts on surface operators and line operators on them. Our description of the 2-categories of surface operators at $t=i$ and $t=1$ can be combined with electric-magnetic duality to give a statement that certain 2-categories attached to the group $U(1)$ are equivalent. This can be viewed as a 2-categorical analogue of the results of \cite{KW}. GL-twisted gauge theory at $t=0$ is particularly interesting. This value of $t$ is preserved by Montonen-Olive duality and, as explained in \cite{Kap:qGL}, the corresponding 4d TFT provides a natural setting for understanding Quantum Geometric Langlands Duality. On the other hand, it was shown in \cite{Kap:qGL} that this TFT does not admit either 't Hooft or Wilson line operators. This presents a problem for the mathematical formulation of the quantum Langlands duality. We will see that while the category of bulk line operators at $t=0$ is indeed rather boring, the abelian theory admits surface operators (not of Gukov-Witten type) whose categories of line operators are quite rich and are acted upon in a nontrivial way by electric-magnetic duality. Hopefully, these observations can be extended to the nonabelian case. Reduction to 3d is useful only insofar as one can understand and classify boundary conditions in 3d TFTs. Up to now, the only 3d TFT where this has been achieved has been the Rozansky-Witten model \cite{KRS,KR}. In this paper we encounter a number of rather unfamiliar 3d TFTs, such a B-type topological gauge theory in 3d and a gauged version of the Rozansky-Witten model. These theories are of independent interest, and we describe some of their properties, including nonlocal observables, in appendices C and D. We will also need to understand categories of branes in certain 2d TFTs obtained by reducing 3d TFTs on an interval. Sometimes these 2d TFTs are topological sigma-models (of type A or B), and then the categories of branes are known (the Fukaya-Floer category and the derived category of coherent sheaves, respectively). In other cases they are gauged topological sigma-models which again can be of type A or B. In appendix A we describe the category of branes for the gauged B-model whose target is a point (i.e. for the B-type gauge theory). In appendix B we describe the category of branes for the gauged B-model with a general target space. Not surprisingly, we find that the category of branes in this TFT is closely related to the equivariant derived category of coherent sheaves and under certain assumptions is equivalent to it. In appendix F we study yet another 2d TFT, the A-type topological gauge theory. We show that when the gauge group is $U(1)$, this model is isomorphic to a B-model whose target is a graded bosonic manifold. This isomorphism allows one to identify the category of branes in the A-type gauge theory This result may be regarded as a physical counterpart of the equivalence between the $U(1)$-equivariant constructible derived category of sheaves over a point and the derived category of modules over the cohomology of the classifying space of $U(1)$ \cite{BL}. A.K. would like to thank R.~ Bezrukavnikov, A.~Braverman, D. ~Orlov, V.~Lunts and especially L.~Rozansky for useful discussions. K.S. acknowledges the support of the Jack Kent Cooke Foundation. This work was supported in part by the DOE grant DE-FG02-92ER40701. \section{Topological field theory, categories and 2-categories: a brief summary} This section is devoted to a brief review of the relationship between 2-categories and Topological Field Theory in two, three, and four dimensions. Readers who are familiar with this subject may skip the section. \subsection{Two-dimensional TFT and categories of branes} For us, a category is a generalization of an algebra, "an algebra with many objects". That is, instead of one vector space $V$ with a multiplication map $V\otimes V\rightarrow V$ we have a set ${\rm Ob}$, a collection of vector spaces $V_{{\mathsf A}{\mathsf B}}$, ${\mathsf A},{\mathsf B}\in {\rm Ob}$, and composition maps $V_{{\mathsf A}{\mathsf B}}\otimes V_{{\mathsf B}{\mathsf C}}\rightarrow V_{{\mathsf A}{\mathsf C}}$. These composition maps must be associative, in an obvious sense. In particular, for each ${\mathsf A}$ the space $V_{{\mathsf A}\A}$ is a (possibly noncommutative) algebra. We will assume in addition that all these algebras have unit elements. The set ${\rm Ob}$ is called the set of objects, and the vector spaces $V_{{\mathsf A}{\mathsf B}}$ are called spaces of morphisms. An element of $V_{{\mathsf A}\A}$ is called an endomorphism of ${\mathsf A}$, and $V_{{\mathsf A}\A}$ is called the endomorphism algebra of $A$. It is common to denote $V_{{\mathsf A}{\mathsf B}}={\rm Mor}({\mathsf A},{\mathsf B})$ and $V_{{\mathsf A}\A}={\rm Mor}({\mathsf A},{\mathsf A})={\rm End}({\mathsf A})$. In physical applications the vector spaces are always complex and usually have integral grading (by some sort of $U(1)$ charge). It is well-known by now that the set of boundary conditions in a 2d TFT has the structure of a category. The set ${\rm Ob}$ of this category is the set of boundary conditions, and the vector space $V_{{\mathsf A}{\mathsf B}}$ is the space of states of the TFT on an (oriented) interval with boundary conditions ${\mathsf A}$ and ${\mathsf B}$. Composition of morphisms arises from the fact that the space $V_{{\mathsf A}{\mathsf B}}$ can be interpreted as the space of local operators sitting at the junction of two segments of the boundary with boundary conditions ${\mathsf A}$ and ${\mathsf B}$ (Fig.1), and from the fact that local operators can be fused together (Fig.2). \begin{figure}[htbp] \label{fig:SOGL_op} \centering \includegraphics[height=2in]{op} \caption{Morphisms in the category of boundary conditions correspond to local operators sitting at the junction of two segments of the boundary.} \end{figure} \begin{figure}[htbp] \label{fig:SOGL_opfusion} \centering \includegraphics[height=2in]{opfusion} \caption{Composition of morphisms is achieved by merging the insertion points of the local operators. We use $\cdot$ to denote this operation.} \end{figure} In the mathematical literature it is common to denote objects by marked points and elements of vector spaces $V_{{\mathsf A}{\mathsf B}}$ by arrows connecting the points. From the physical viewpoint it is more natural to denote objects by marked segments of an oriented line, and morphisms by points sitting at the junction of two consecutive segments. Let us recall two simple examples of 2d TFTs which will be important for us. The first one is a B-model with a target $X$, where $X$ is a complex manifold with a holomorphic volume form (i.e. a possibly noncompact Calabi-Yau manifold). The corresponding category of boundary conditions has been argued to be equivalent to the bounded derived category of coherent sheaves on $X$, which is denoted $D^b(Coh(X))$. Its objects can be thought of as complexes of holomorphic vector bundles on $X$ or complex submanifolds of $X$. The second one is an A-model with target $Y$, where $Y$ is a symplectic manifold. The corresponding category of boundary conditions is believed to be equivalent to a version of the Fukaya-Floer category. Its simplest objects are Lagrangian submanifolds of $Y$ equipped with unitary vector bundles with flat connections. In the 2d context one usually refers to boundary conditions as branes and talks about A-branes and B-branes. A and B-models do not exhaust the possibilities even in two dimensions. Below we will encounter other, less familiar, 2d TFTs and their categories of branes. \subsection{Two-dimensional TFTs and 2-categories} A boundary of a 2d TFT can be regarded as a boundary between a nontrivial TFT and a trivial TFT. More generally, one can consider boundaries between arbitrary pairs of 2d TFTs. Such boundaries may be called defect lines, or walls. The set of all walls between a fixed pair of TFTs has the structure of a category. To see this, let ${\mathbb X}$ and ${\mathbb Y}$ denote our chosen pair of TFTs, and let $\bar{\mathbb X}$ denote the theory ${\mathbb X}$ with a reversed parity. By folding back the worldsheet at the wall location (see Fig. 3), we see that a wall between ${\mathbb X}$ and ${\mathbb Y}$ is the same as a boundary of the theory $\bar{\mathbb X}\times{\mathbb Y}$. \begin{figure}[htbp] \label{fig:SOGL_foldingtrick} \centering \includegraphics[height=2in]{foldingtrick} \caption{A wall separating theories ${\mathbb X}$ and ${\mathbb Y}$ is equivalent to a boundary of theory $\bar{\mathbb X}\times{\mathbb Y}$.} \end{figure} Thus we may appeal to the previous discussion and conclude that walls are objects of a category ${\mathbb V}_{{\mathbb X}{\mathbb Y}}$. Given any two walls ${\mathsf A},{\mathsf B}\in{\rm Ob}({\mathbb V}_{{\mathbb X}{\mathbb Y}})$, the space of morphisms from ${\mathsf A}$ to ${\mathsf B}$ is the space of local operators which can be inserted at the junction of ${\mathsf A}$ and ${\mathsf B}$. Composition of morphisms is obtained by fusing local operators sitting on a defect line. There is an obvious ``fusion'' operation on the set of walls: given a wall between theories ${\mathbb X}$ and ${\mathbb Y}$ and a wall between theories ${\mathbb Y}$ and $\mathcal Y$ we may fuse them and get a wall between theories ${\mathbb X}$ and $\mathcal Y$ (Fig. 4). One can describe the situation mathematically by saying that the set of 2d TFTs has the structure of a 2-category. A 2-category has objects, morphisms, and 2-morphisms (morphisms between morphisms). In the present case, objects are 2d TFTs. The set of morphisms from an object ${\mathbb X}$ to an object ${\mathbb Y}$ is the set of walls between theories ${\mathbb X}$ and ${\mathbb Y}$. Fusion of walls gives rise to a way of composing morphisms. Given any two walls between the same pair of TFTs, the space of 2-morphisms between them is the space of local operators which can be inserted at the junction of these two walls. One can put this slightly differently and say that a 2-category has a collection of objects (which are 2d TFTs in our case), and for any pair of objects ${\mathbb X}$ and ${\mathbb Y}$ one has a category of morphisms ${\mathbb V}_{{\mathbb X}{\mathbb Y}}$ (which is the category of walls in our case). Fusion of walls means that there is a way to ``compose'' categories of morphisms. That is, given an object ${\mathsf A}$ of the category ${\mathbb V}_{{\mathbb X}{\mathbb Y}}$ and an object ${\mathsf B}$ of the category ${\mathbb V}_{{\mathbb Y}\mathcal Y}$ there is a rule which determines an object ${\mathsf A}\otimes {\mathsf B}$ of the category ${\mathbb V}_{{\mathbb X}\mathcal Y}$. \begin{figure}[htbp] \label{fig:SOGL_wallfusion} \centering \includegraphics[height=2in]{wallfusion} \caption{1-morphisms of the 2-category of 2d TFTs correspond to walls, and composition of 1-morphisms corresponds to fusing walls. This operation is denoted $\otimes$.} \end{figure} This is not all though: one can fuse not only walls, but walls with local operators inserted on them (Fig. 5). This determines composition maps on 2-morphisms. \begin{figure}[htbp] \label{fig:SOGL_opwall} \centering \includegraphics[height=2in]{opwall} \caption{Composition of 2-morphisms of the 2-category of 2d TFTs is achieved by fusing the walls on which they are inserted. The corresponding operation is denoted $\otimes$.} \end{figure} This new composition is different from the composition of local operators regarded as morphisms in the category ${\mathbb V}_{{\mathbb X}{\mathbb Y}}$. The composition maps enjoy various properties which can be deduced by staring at the pictures of fusing walls and making use of the metric independence. For example, the old and new compositions commute, as illustrated in Fig. 6. \begin{figure}[htbp] \label{fig:SOGL_commutative} \centering \includegraphics[height=3.5in]{commutative} \caption{Local operators inserted on walls may be regarded either as 2-morphisms of the category of 2d TFTs, in which case composition corresponds to fusing ``horizontally," or they may be regarded as morphisms of the category of boundary conditions, in which case composition corresponds to fusing ``vertically". These two operations commute. } \end{figure} Even if one is interested in a particular TFT, the notion of a 2-category is useful. Namely, defect lines in a 2d TFT form a 2-category with a single object. In this case there is only one category of morphisms, ${\mathbb V}_{{\mathbb X}\X}$, with additional structure coming from the fact that fusing two defect lines gives another defect line in the same theory. This structure allows one to define a rule for ``tensoring'' objects of ${\mathbb V}_{{\mathbb X}\X}$: $$ ({\mathsf A},{\mathsf B})\mapsto {\mathsf A}\otimes {\mathsf B}\in {\rm Ob}({\mathbb V}_{{\mathbb X}\X}) $$ and morphisms $$ {\rm Mor}({\mathsf A},{\mathsf B})\otimes {\rm Mor}({\mathsf C},{\mathsf D})\rightarrow {\rm Mor}({\mathsf A}\otimes {\mathsf C},{\mathsf B}\otimes {\mathsf D}) $$ This tensoring does not need to be commutative, in general. A category with such additional ``tensor'' structure is called a monoidal category. It should be clear from the above that a monoidal category is the same thing as a 2-category with a single object. Among all defect lines in a given 2d TFT there is a trivial defect line $\1$ which is equivalent to no defect at all. We may call it the invisible defect line. It is the identity object in the monoidal category of defect lines, in the sense that fusing it with any other defect line ${\mathsf A}$ gives back ${\mathsf A}$. Endomorphisms of the trivial defect line (i.e. elements of the vector space ${\rm Mor}(\1,\1)$) are the same as local operators in the bulk. The simplest example of a monoidal category is the category of vector spaces, with the usual tensor product. It can be regarded as the 2-category of defect lines in a trivial 2d TFT (say, a topological sigma-model whose target is a point). Defect lines in Landau-Ginzburg TFTs have been studied in \cite{LGdefects1, LGdefects2}. One can fuse a defect line in a given 2d TFT with any boundary condition and get a new boundary condition in the same TFT. This defines an ``action'' of the monoidal category of defect lines on the category of branes. Mathematically this can be described using the notion of a module category. Since a monoidal cateory is a categorification of the notion of an algebra, it is natural to define a module category over a monoidal category as a categorification of the notion of a module over an algebra. A definition of a module category ${\mathbb W}$ over a monoidal category ${\mathbb V}$ involves a rule for ``multiplying'' an object on ${\mathbb W}$ by an object of ${\mathbb V}$: $$ ({\mathsf A},{\mathsf C})\mapsto {\mathsf A}\cdot {\mathsf C}\in{\rm Ob}({\mathbb W}),\quad \forall {\mathsf A}\in {\rm Ob}({\mathbb W}),\ \forall {\mathsf C}\in{\rm Ob}({\mathbb V}), $$ as well as a rule for multiplying morphisms, i.e. a map $$ {\rm Mor}_{\mathbb W}({\mathsf A},{\mathsf B})\otimes {\rm Mor}_{\mathbb V}({\mathsf C},{\mathsf D})\rightarrow {\rm Mor}_{\mathbb W} ({\mathsf A}\cdot {\mathsf C},{\mathsf B}\cdot{\mathsf D}). $$ The latter rule encodes the fact that we can fuse a junction of two defect lines with a junction of two boundary conditions and get a new junction of two new boundary conditions. An important idea which we systematically use in this paper is that some properties of codimension-2 defects can be studied using dimensional reduction. We have already seen a simple example of this: the space of local operators sitting at the junction of two boundary conditions ${\mathsf A}$ and ${\mathsf B}$ can be thought of as the space of states of a 1d field theory (i.e. quantum mechanics) obtained by compactifying the 2d TFT on an interval with boundary conditions ${\mathsf A}$ and ${\mathsf B}$. Another example is the space of local operators in the bulk. It is well known that it can be identified with the space of states of a 1d field theory obtained compactifying the 2d TFT on a circle. The argument is essentially the same in both cases. After one excises a tubular neigborhood of the local operator, the operator insertion is replaced by a boundary whose collar neighborhood looks like ${\mathbb R}_+\times I$ in the first case and ${\mathbb R}_+\times S^1$ in the second case. Then one uses the fact that in a TFT the size of the tubular neighborhood does not matter, and one can regard any boundary condition on the newly created boundary as a local operator. It is important to note that this reinterpretation of codimension-2 defects in terms of a lower-dimensional theory causes ``information loss''. For example, we cannot compute the composition $V_{{\mathsf A}{\mathsf B}}\otimes V_{{\mathsf B}{\mathsf C}}\rightarrow V_{{\mathsf A}{\mathsf C}}$ if we view the vector spaces involved as spaces of states of three 1d field theories. Similarly, we cannot see the commutative algebra structure on the space of bulk local operators if we view it as the space of states of a 1d field theory. \subsection{Three-dimensional TFT and 2-categories of boundary conditions} When we move to dimension three, we find that boundary conditions in a 3d TFT also form a 2-category. To see this, let us first consider a trivial 3d TFT (say, a 3d topological sigma-model whose target is a point). Even though there are no bulk degrees of freedom in this case, we may consider putting any 2d TFT on the boundary. Thus the set of all boundary conditions for the trivial 3d TFT is the set of all 2d TFTs, which form a 2-category. Walls between 2d TFTs now can be interpreted as defect lines on the 2d boundary of a 3d worldvolume. The monoidal category of defect lines for a given 2d TFT can be reinterpreted as the monoidal category of boundary line operators for a particular boundary condition. If the 3d TFT in the bulk is nontrivial, we can still couple it to a 2d TFT on the boundary. Different boundary conditions are distinguished by the type of 2d TFT on the boundary and by its coupling to the bulk degrees of freedom. For a concrete example of how this works in the Rozansky-Witten 3d TFT, see \cite{KRS}. Again one may consider boundary defect lines separating different boundary conditions, and their fusion and fusion of local operators on boundary defect lines can be described by a 2-category structure on the set of boundary conditions. If we focus on a particular boundary condition, then the set of defect lines on this boundary has the structure of a monoidal category. Mimicking what we did in 2d, we may consider walls, or surface operators, between different 3d TFTs. The set of walls between any two 3d TFTs ${\mathfrak K}$ and ${\mathfrak L}$ has the structure of a 2-category. One way to see it is to fold the worldvolume along the defect surface and reinterpret the wall as a boundary condition for the theory $\bar{\mathfrak K}\times{\mathfrak L}$, where $\bar{\mathfrak K}$ is the parity-reversal of the theory ${\mathfrak K}$. Furthermore, just like in 2d, we can fuse walls with defect lines and local operators on them (Fig. 7). \begin{figure}[htbp] \label{fig:SOGL_surfacefusion} \centering \includegraphics[height=1.8in]{surfacefusion} \caption{Regarding defect lines as 2-morphisms and local operators as 3-morphisms of the 3-category of 3d TFTs gives rise to yet another composition operation between them, which we denote $\odot$. } \end{figure} Altogether, one can summarize the situation by saying that 3d TFTs form a 3-category. Its objects are 3d TFTs, its morphisms are walls between 3d TFTs, its 2-morphisms (morphisms between morphisms) are defect lines on walls, and its 3-morphisms (morphisms between 2-morphisms) are local operators sitting on defect lines. If we restrict attention to surface operators in a particular 3d TFT, we get a 3-category with a single object. Equivalently, the 2-category of surface operators in a 3d TFT has an extra structure which allows one to fuse objects, morphisms and 2-morphisms. In other words, it is a monoidal 2-category. It has an identity object (the trivial surface operator) whose endomorphisms can be thought of as defect lines in the bulk. To determine the category of bulk defect lines in a given 3d TFT one may use the dimensional reduction trick. To apply it, we excise a tubular neigborhood of a defect line and replace the defect line by a suitable boundary condition on its boundary. The collar neighborhood of the newly created boundary locally looks like $S^1\times{\mathbb R}_+\times {\mathbb R}$, where ${\mathbb R}_+$ corresponds to the ``radial'' direction. Therefore we may identify the defect line with the boundary condition in the 2d TFT obtained by compactifying the 3d TFT on a circle. This trick allows one to determine the category of bulk defect lines by studying the category of branes in a 2d TFT. We lose some information in this way: the category of bulk defect lines in a 3d TFT is in fact a braided monoidal category (i.e. a category with a quasi-commutative tensor product), but this structure cannot be seen from the 2d viewpoint. Similarly, the study of categories of boundary defect lines reduces to the study of the category of branes in a 2d TFT obtained by compactifying the 3d TFT on an interval. If the boundary defect line separates two boundary conditions ${\mathbb X}$ and ${\mathbb Y}$, then the boundary conditions on the endpoints of the interval should be ${\mathbb X}$ and ${\mathbb Y}$. This 2d viewpoint entails some information loss: for example, it does not allow one to compute the monoidal structure on the category ${\mathbb V}_{{\mathbb X}\X}$. \subsection{Four-dimensional TFT and 2-categories of surface operators} Boundary conditions in a 4d TFT form a 3-category. For example, if the 4d TFT is trivial, its 3-category of boundary conditions is the 3-category of all 3d TFTs. Similarly, walls (codimension-1 defects) in a given 4d TFT form a monoidal 3-category. In this paper we will avoid dealing with such complicated structures and will focus instead on defects of codimension two, i.e. surface operators in a 4d TFT. Such surface operators form a 2-category. One way to see it is to apply the dimensional reduction trick: excise a tubular neighborhood of a surface operator and replace the operator by a suitable boundary condition on the newly created boundary. The collar neighborhood of the boundary looks locally like $S^1\times {\mathbb R}_+\times{\mathbb R}^2$, so we may reinterpret a surface operator as a boundary condition in a 3d TFT obtained by compactifying the 4d theory on a circle. Then we can appeal to the known fact that boundary conditions in a 3d TFT form a 2-category. Of course, one can also explain the meaning of this 2-category structure directly in 4d terms. The type of a surface operator may jump across a defect line, and one can regard defect lines on surface operators as morphisms in a 2-category. Local operators sitting at a junction point of two surface defect lines are 2-morphisms. The 4d viewpoint also makes it clear that the 2-category of surface operators has a rich extra structure. First of all, one may fuse surface operators together with defect lines and local operators sitting on them. This gives rise to a monoidal structure on the 2-category of surface operators. Second, by moving surface operators around one can easily see that the fusion operation is quasi-commutative, i.e. one gets a braided monoidal 2-category.\footnote{A possible mathematical definition of a braided monoidal 2-category is spelled out in \cite{KapVoe}. However, it appears that braided monoidal structures which arise in 4d TFT are of a rather special kind. In particular, the braiding is always invertible.} In this paper we will use the dimensional reduction trick to study the 2-category of surface operators and leave the understanding of the braided monoidal structure on this 2-cateory for future work. \section{Review of the GL-twisted theory} The bosonic fields in the GL-twisted 4d theory are a gauge field $A_\mu$ (a connection on a principal $G$-bundle $\P$ over a 4-manifold $M_4$), a 1-form $\phi_\mu dx^\mu$ with values in ${\rm Ad}(\P)$, and a 0-form $\sigma$ with values in the complexification of ${\rm Ad}(\P)$. The conventions are the same as in \cite{KW}; in particular, real adjoint-valued fields are regarded as anti-Hermitian, and the covariant derivative in the adjoint representation takes the form $d_A=d+[A,\cdot]$. The fermionic fields are a pair of ${\rm Ad}(\P)_{\mathbb C}$-valued 1-forms $\psi$ and ${\tilde\psi}$, a pair of ${\rm Ad}(\P)_{\mathbb C}$-valued 0-forms $\eta$ and ${\tilde\eta}$, and an ${\rm Ad}(\P)_{\mathbb C}$-valued 2-form $\chi$. The fields $A$ and $\phi$ have ghost number $0$, the fields $\psi$ and ${\tilde\psi}$ have ghost number $1$, the fields $\eta,{\tilde\eta},$ and $\chi$ have ghost number $-1$, and the field $\sigma$ has ghost number $2$. The BRST transformations are \begin{align*} \delta A &=i(\psi+t{\tilde\psi}),\\ \delta\phi & =i(t\psi-{\tilde\psi}),\\ \delta\sigma &=0,\\ \delta{\bar\sigma} &=i(\eta+t {\tilde\eta}),\\ \delta \psi & = d_A\sigma+t[\phi,\sigma]\\ \delta{\tilde\psi} & = t d_A\sigma -[\phi,\sigma],\\ \delta\eta & = t d_A^*\phi+[{\bar\sigma},\sigma],\\ \delta{\tilde\eta} & = - d_A^*\phi+t[{\bar\sigma},\sigma],\\ \delta\chi & =\frac{1+t}{2}(F-\frac12 [\phi,\phi]+ * d_A\phi)+\frac{1-t}{2}(* (F-\frac12 [\phi,\phi])-d_A\phi). \end{align*} Here $t$ takes values in ${\mathbb C}\bigcup \{\infty\}$, $\bar\sigma=-\sigma^\dagger$, $*$ is the 4d Hodge star operator, and $d^*=* d *$. For $t\neq \pm i$ the action can be wriiten as a BRST-exact term plus a term which depends only on the topology of the bundle $\P$: \begin{equation} S=\delta \int_{M_4} V-\frac{\Psi}{4\pi i}\int_{M_4} {\rm Tr}\, F\wedge F, \end{equation} where $$ \Psi=\frac{\theta}{2\pi}+\frac{4\pi i}{e^2}\frac{t^2-1}{t^2+1}. $$ Here $\theta$ is the theta-angle of the 4d gauge theory and $e^2$ is the gauge coupling. The explicit form of $V$ can be found in \cite{KW}. The simplest surface operators have been introduced by Gukov and Witten \cite{GW}. They are disorder operators corresponding to a codimension-2 singularity in the fields of the form $$ A=\alpha d\theta, \quad \phi=\beta \frac{dr}{r}-\gamma d\theta. $$ Here $\alpha$ is an element of a maximal torus ${\mathsf T}$ of $G$, and $\beta,\gamma$ are elements of the Lie algebra ${\mathfrak t}$ of ${\mathsf T}$. For simplicity, let us assume that the triple $(\alpha,\beta,\gamma)$ breaks $G$ down to ${\mathsf T}$. Gauge transformations which preserve ${\mathsf T}$ form the Weyl group ${\mathcal W}$; the triplet $(\alpha,\beta,\gamma)$ is defined up to the action of ${\mathcal W}$ on ${\mathsf T} \times {\mathfrak t}\times{\mathfrak t}$. All fields other than $A$ and $\phi$ are nonsingular. The surface operator depends on an additional parameter $\eta$ taking values in the torus ${\rm Hom}(\Lambda_{\rm cochar},U(1))$. Here $\Lambda_{\rm cochar}$ is the lattice of magnetic charges ${\rm Hom}(U(1),{\mathsf T})$. Equivalently, as explained in \cite{GW}, $\eta$ can be thought of as taking values in ${{}^L\mathsf T}$, the maximal torus of the Langlands-dual group. The parameter $\eta$ arises as follows. First, note that the above singularity in the fields breaks the gauge group down to ${\mathsf T}$. Thus if $D$ is the codimension-2 submanifold on which the surface operator is supported, the restriction of the gauge field to $D$ has a first Chern class $c_1\vert_D$ taking values in $\Lambda_{\rm cochar}$. Given $\eta$ we can insert into the path-integral a phase factor $$ \eta(c_1(D)). $$ This factor depends only on the behavior of the gauge field on $D$ and can be regarded as an $\eta$-dependent modification of the surface operator defined above. Gukov-Witten surface operators are BRST-invariant for arbitrary $t$, but their properties depend on $t$. We will see below that there are many other surface operators. In what follows we will focus on the cases $t=i$, $t=1$, and $t=0$. The first two cases are exchanged by S-duality (at zero $\theta$-angle) and play a prominent role in the physical approach to the Geometric Langlands Program \cite{KW}. The last case is self-dual and is the most natural starting point for understanding Quantum Geometric Langlands Duality \cite{Kap:qGL}. \section{Surface operators at $t=i$: the abelian case} \subsection{Reduction to 3d} As explained in \cite{GW}, at $t=i$ varying the parameters $\beta$ and $\eta$ changes the surface operator only by BRST-exact terms. Thus Gukov-Witten operators depend on a single complex parameter $\alpha-i\gamma$. But there exist much more general surface operators. To study them systematically, it is convenient to use the fact that surface operators in the 4d TFT are in 1-1 correspondence with boundary conditions in the 3d TFT compactified on a circle. The advantage of the 3d viewpoint is that the problem of classification of boundary conditions is more familiar. In particular, for $t=i$ the 3d TFT that one gets is a gauged version of the Rozansky-Witten model, so we can use many of the results of \cite{KRS} where boundary conditions for the Rozansky-Witten model have been studied. In this section we consider the case $G=U(1)$. Reduction to 3d amounts to declaring all fields to be independent of the $x^4$ direction which is periodic with period $2\pi$. The reduced theory has the following bosonic fields: a 3d gauge field $A$, a 1-form $\phi$, a complex 0-form $\sigma$, and a pair of 0-forms $A_4$ and $\phi_4$. More properly, one should work with a $U(1)$-valued scalar $\exp(-2\pi A_4)$ which represents the holonomy of the gauge field along the compact direction. This field is invariant with respect to $x^4$-dependent gauge transformations $$ A_4\mapsto A_4 + im,\quad m\in{\mathbb Z}. $$ The fermionic fields are 1-forms $\psi,{\tilde\psi},\chi,{\tilde\chi}$, and 0-forms $\eta,{\tilde\eta},\psi_4,{\tilde\psi}_4$. At $t=i$ it is convenient to combine $A_4$ and $\phi_4$ into a complex 0-form $\tau=A_4+i\phi_4$, or more properly into a gauge-invariant ${\mathbb C}^*$-valued scalar $\exp(-2\pi\tau)$. Then $\tau$ and $\sigma$ are BRST-invariant. We also define the complex 3d gauge field ${\mathcal A}=A+i\phi$ which is BRST-invariant and the corresponding curvature ${\mathcal F}=d{\mathcal A}$. The BRST transformations of other fields are \begin{align*} \delta (A-i\phi) &=2i(\psi+i{\tilde\psi}),\\ \delta{\bar\sigma} &=i(\eta+i {\tilde\eta}),\\ \delta \tau & =-2i(\psi_4+i{\tilde\psi}_4) ,\\ \delta \psi & = d\sigma\\ \delta{\tilde\psi} & = i d\sigma,\\ \delta\psi_4 &= 0,\\ \delta{\tilde\psi}_4 &=0,\\ \delta\eta & = i d^\star\phi,\\ \delta{\tilde\eta} & = - d^\star\phi,\\ \delta\chi & = {\mathcal F},\\ \delta{\tilde\chi}& =d\tau. \end{align*} Here $d^\star=\star d\star$ and $\star$ denotes the 3d Hodge star operator. These BRST-transformations are nilpotent off-shell. One can make them nilpotent on-shell by introducing a suitable auxiliary field, as discussed in the appendix E. The 3d action contains both a BRST-exact metric-dependent term and a BRST-closed metric-independent term. Its explicit form is given in the appendix E. The analysis of boundary conditions in the 3d theory is greatly facilitated by the observation that this 3d theory decomposes into two independent sectors, the Rozansky-Witten model with target $T^*{\mathbb C}^*$ and a topological $U(1)$ gauge theory. Let us discuss these two 3d TFTs in turn. \subsection{Rozansky-Witten model with target $T^*{\mathbb C}^*$} The fields of this model are a subset of the fields of the 3d theory listed above. The bosonic ones are the ${\mathbb C}^*$-valued scalar $h=\exp(-2\pi \tau)$ and the ${\mathbb C}$-valued scalar $\sigma$. The fermionic ones are the 0-forms $\psi_4+i{\tilde\psi}_4,\eta+i{\tilde\eta}$ and the 1-forms $\psi-i{\tilde\psi},{\tilde\chi}$. The RW model can be defined for any complex symplectic target space $X$, and $T^*{\mathbb C}^*$ is a special case with the symplectic form $d\tau\wedge d\sigma$. It is shown in appendix E that the correct 3d action arises from the 4d action of the GL-twisted theory upon reduction. For a general $X$ the RW model has ${\mathbb Z}_2$ ghost number symmetry, but as explained in \cite{KRS} when $X$ is a cotangent bundle one can promote it to a $U(1)$ ghost number symmetry by letting the fiber coordinates have ghost number two. This agrees with the fact that $\sigma$ has ghost number two already in the 4d theory. To emphasize that the fiber coordinate has ghost number two we will denote the target manifold $T^*[2]{\mathbb C}^*$. According to \cite{KRS} the simplest boundary conditions in the RW model correspond to complex Lagrangian submanifolds of $X$. If we want to preserve ghost number symmetry, these Lagrangian submanifolds must be invariant with respect to the rescaling $\sigma\mapsto \lambda^2\sigma$, $\lambda\in{\mathbb C}^*$. This requires the Lagrangian submanifold of $T^*{\mathbb C}^*$ to be the conormal bundle of a complex submanifold in ${\mathbb C}^*$. This means that a (closed) ${\mathbb C}^*$-invariant complex Lagrangian submanifold is either the zero section $\sigma=0$ or one of the fibers of the cotangent bundle given by $\tau=\tau_0$. The zero section boundary condition plays a special role and will be denoted ${\mathbb X}_0$ in this subsection. More general boundary conditions correspond to families of B-models or Landau-Ginzburg models parameterized by points in a complex Lagrangian submanifold. As mentioned in \cite{KRS} and explained in more detail in \cite{KR} it is sufficient to restrict oneself to the case when the Lagrangian submanifold is the zero section $\sigma=0$. One can describe these boundary conditions more algebraically as follows. Recall that the category of boundary line operators on the boundary ${\mathbb X}_0$ is a monoidal category which we denote ${\mathbb V}_{{\mathbb X}_0{\mathbb X}_0}$. Given any boundary condition ${\mathbb X}$ one may consider the category ${\mathbb V}_{{\mathbb X}\X_0}$ of boundary defect lines which may separate ${\mathbb X}$ from ${\mathbb X}_0$. This category is a module category over the monoidal category ${\mathbb V}_{{\mathbb X}_0{\mathbb X}_0}$. It was proposed in \cite{KRS} that this module category completely characterizes the boundary condition ${\mathbb X}$. Concretely, in the case of the RW model with target $T^*[2]{\mathbb C}^*$ the category of boundary line operators ${\mathbb V}_{{\mathbb X}_0{\mathbb X}_0}$ is equivalent to $D^b(Coh({\mathbb C}^*))$. One way to see it is to reduce the RW model on an interval with the boundary condition ${\mathbb X}_0$ on both boundaries. The resulting 2d TFT is a B-model with target ${\mathbb C}^*$, and its category of branes may be identified with $D^b(Coh({\mathbb C}^*))$. The 2d viewpoint does not allow one to determine the monoidal structure, but one can show that it is given by the usual derived tensor product \cite{KRS,KR}. It was further argued in \cite{KRS,KR} that the 2-category of boundary conditions for the RW model with target $T^*[2]{\mathbb C}^*$ is equivalent to the derived 2-category of module categories over $D^b(Coh({\mathbb C}^*))$. That is, it is the 2-category of derived categorical sheaves over ${\mathbb C}^*$ as defined by B.~Toen and G.~Vezzosi \cite{ToVe}. This provides an algebraic description of boundary line operators and their OPEs for all boundary conditions. \subsection{B-type topological gauge theory with gauge group $U(1)$ } There are two different topological gauge theories in 3d which can be obtained by twisting $N=4$ $d=3$ super-Yang-Mills theory. The first one is the dimensional reduction of the Donaldson-Witten twist of $N=2$ $d=4$ super-Yang-Mills theory. The second one is intrinsic to 3d and has been first discussed by Blau and Thompson \cite{BT}. We will refer to them as A-type and B-type topological gauge theories respectively. The reason for this terminology is that the BPS equations in the former theory are elliptic, as in the usual A-model, while in the latter theory they are overdetermined, as in the usual B-model. The definition and some properties of the B-type 3d gauge theory (for a general gauge group) are described in the appendix C. In this subsection we only deal with the abelian case. Consider the 3d bosonic fields $A$, $\phi$ and the fermionic fields $\psi+i{\tilde\psi},\chi,\eta-i{\tilde\eta},\psi_4-i{\tilde\psi}_4$. It is easy to check that their BRST transformations at $t=i$ are exactly the same as for the B-type 3d gauge theory. The action has a BRST-exact metric-dependent piece and a BRST-closed metric-independent piece: $$ S=-\frac{1}{2e^2}\, \delta\int_{M_3} \left(\chi\wedge \star {\mathcal F}-\frac{i}{2}(\eta-i{\tilde\eta})\wedge \star d^\star\phi\right)+\frac{1}{2 e^2}\int_{M_3}(\psi_4-i{\tilde\psi}_4) d\chi $$ In principle we should gauge-fix the theory and modify the BRST operator appropriately; we leave this as an exercise for the reader. As in any gauge theory, the most natural boundary conditions are the Dirichlet and Neumann ones. The Dirichlet condition requires the restriction of $A+i\phi$ to the boundary to be trivial. In addition, one requires $\phi_3$ (the component of $\phi$ orthogonal to the boundary) to satisfy the Neumann condition $\partial_3\phi_3=0$. BRST-invariance then fixes the boundary conditions for fermions: the restriction of the forms $\psi+i{\tilde\psi}$, $\chi$ and $\eta-i{\tilde\eta}$ to the boundary must vanish, The Neumann boundary condition leaves the restriction of ${\mathcal A}$ to the boundary unconstrained but requires the restriction of the 1-form $\star{\mathcal F}=\star d{\mathcal A}$ to vanish. In addition $\phi_3$ must satisfy the Dirichlet boundary condition, i.e. it must take a prescribed value on the boundary. In the Neumann case BRST-invariance requires the restrictions of the fermions $\star\chi$, $\psi_3+i{\tilde\psi}_3$ and $\psi_4-i{\tilde\psi}_4$ to vanish. Note that in the Dirichlet case the gauge group is completely broken at the boundary, while in the Neumann case it is unbroken. The Dirichlet condition does not have any parameters, while the Neumann condition seems to depend on a single real parameter $\beta$, the boundary value of $\phi_3$. On the quantum level there is another parameter: we can add to the action a boundary topological term $$ \theta \int_{\partial M_3} \frac{{\mathcal F}}{2\pi} $$ In fact, both parameters are irrelevant, in the sense that topological correlators do not depend on them. The irrelevance of the parameter $\theta$ follows from the fact that the above topological term is BRST-exact and equal to $$ \frac{\theta}{2\pi}\delta \int_{\partial M_3} \chi $$ To see the irrelevance of the parameter $\beta$, note that to shift $\beta$ we need to add to the action a boundary term proportional to $$ \int_{\partial M_3} \partial_3\phi_3 $$ Since $\phi_1$ and $\phi_2$ vanish on the boundary, this is also BRST-exact and proportional to $$ \delta \int_{\partial M_3} (\eta-i{\tilde\eta}). $$ Following the same line of thought as in \cite{KRS}, one can try to describe the 2-category of boundary conditions in this theory by picking a distinguished boundary condition ${\mathbb X}_0$ and characterizing any other boundary condition ${\mathbb X}$ by the category ${\mathbb V}_{{\mathbb X}\X_0}$ of defect line operators between ${\mathbb X}$ and ${\mathbb X}_0$. That is, one attaches to any boundary condition ${\mathbb X}$ a module category ${\mathbb V}_{{\mathbb X}\X_0}$ over the monoidal category ${\mathbb V}_{{\mathbb X}_0{\mathbb X}_0}$. An obvious guess for the distinguished boundary condition is the free (Neumann) one since it leaves the gauge group unbroken. To determine the category of boundary line operators ${\mathbb V}_{{\mathbb X}_0{\mathbb X}_0}$ for this boundary condition, one may reduce the 3d theory on an interval and study the category of branes in the resulting 2d TFT. In the Neumann case, reduction on an interval gives the following result: the bosonic fields are the gauge field $A$ and the 1-form $\phi$, the fermionic ones are the 0-form $\eta-i{\tilde\eta}$, the 1-form $\psi+i{\tilde\psi}$ and the 2-form $\chi$. This is the field content of a B-type topological gauge theory in 2d, see appendix A. It is easy to check that the BRST transformations of these fields are also the same as in the B-type 2d gauge theory. The category of branes for this 2d TFT is the category of graded finite-dimensional representations of $G=U(1)$, see appendix A for details. This is because the only boundary degrees of freedom one can attach are described by a vector space which carries a representation of the gauge group. The monoidal structure cannot be determined from 2d considerations, but it easy to see that it is given by the usual tensor product. Indeed, as described in the appendix, a brane corresponding to a representation space $V$ is obtained by inserting the holonomy of the complex connection ${\mathcal A}=A+i\phi$ in the representation $V$ into the path-integral. From the 3d viewpoint this means that the corresponding boundary line operator is the Wilson line operator for ${\mathcal A}$ in the representation $V$. On the classical level, the fusion of two Wilson line operators in representations $V_1$ and $V_2$ gives the Wilson line in representation $V_1\otimes V_2$, and clearly there can be no quantum corrections to this result (the gauge coupling $e^2$ is an irrelevant parameter). To summarize, the monoidal category ${\mathbb V}_{{\mathbb X}_0{\mathbb X}_0}$ is the category of graded finite-dimensional representations of ${\mathbb C}^*$, or equivalently the equivariant derived category of coherent sheaves over a point which we denote $D^b_{{\mathbb C}^*}(Coh(\bullet))$. We propose that the 2-category of boundary conditions is equivalent to the 2-category of module categories over $D^b_{{\mathbb C}^*}(Coh(\bullet))$. To give a concrete class of examples of such a module category, consider a Calabi-Yau manifold $Y$ with a ${\mathbb C}^*$ action. The corresponding B-model can be coupled to the boundary gauge field and provides a natural set of topological boundary degrees of freedom for the 3d gauge theory. The corresponding category of boundary-changing line operators is the ${\mathbb C}^*$-equivariant bounded derived category of $Y$ which is obviously a module category over $D^b_{{\mathbb C}^*}(Coh(\bullet))$. \subsection{Putting the sectors together} It is fairly obvious how to combine the two models. The most basic boundary condition in the full theory is $\sigma=0$ in the RW sector and the free (Neumann) condition in the gauge sector. We will call this the distinguished boundary condition. The bosonic fields which are free on the boundary are the ${\mathbb C}^*$-valued scalar $h=\exp(-2\pi\tau)$ and the restriction of the complex gauge field ${\mathcal A}=A+i\phi$. More general boundary conditions involve a boundary B-model or a boundary Landau-Ginzburg model fibered over ${\mathbb C}^*$ and admitting a ${\mathbb C}^*$-action. The fibration over ${\mathbb C}^*$ determines the coupling to the boundary value of $\tau$, while the ${\mathbb C}^*$-action determines the coupling to the boundary gauge field ${\mathcal A}$. As in the RW model, we can give a more algebraic definition of the set of all boundary conditions in the full theory. This description is also useful because it suggests how to define the 2-category structure of the set of boundary condtions. We consider the monoidal category of boundary line operators for the distinguished boundary condition. This is the category of branes for the 2d TFT obtained by reducing the gauged RW model on an interval. Since the reduction of the B-type 3d gauge theory gives the B-type 2d gauge theory, and the reduction of the RW model gives the B-model with target ${\mathbb C}^*$, the effective 2d TFT is the gauged B-model with target ${\mathbb C}^*$, where the gauge group $U(1)$ acts trivially. As described in appendix B, the corresponding category of branes is equivalent to $D^b_{{\mathbb C}^*}(Coh({\mathbb C}^*))$. The monoidal structure cannot be determined from the 2d considerations, but it is easy to see (given the results for the RW model and the B-type gauge theory in 3d) that it is given by the derived tensor product. Every boundary condition gives rise to a module category over this monoidal category. It is natural to conjecture that the converse is also true, i.e. every reasonable module category over this monoidal category can be thought of as a boundary condition for the full 3d TFT. For example, we may consider a family of Calabi-Yau manifolds parameterized by points of ${\mathbb C}^*$ such that each model in the family has a ${\mathbb C}^*$ symmetry. The corresponding module category is the ${\mathbb C}^*$-equivariant derived category of the total space of the fibration. This gives us a conjectural description of the 2-category of surface operators in the parent 4d gauge theory. Let us describe how Gukov-Witten surface operators fit into this picture. Such operators depend on a complex parameter $h_0=\exp(-2\pi (\alpha-i\gamma))$ taking values in ${\mathbb C}^*$. From the 3d viewpoint, $h_0$ determines the boundary value of the scalar $h=\exp(-2\pi\tau)$ in the RW sector. The other scalar $\sigma$ is left free. Thus the boundary conditions for the RW sector correspond to a Lagrangian submanifold of $T^*[2]{\mathbb C}^*$ given by $h=h_0$ (the fiber over the point $h_0$). The gauge sector boundary conditions are of Neumann type and have no nontrivial parameters.There are no boundary degrees of freedom. From our algebraic viewpoint we may describe this as follows. In the usual RW theory the fiber over $h=h_0$ corresponds to a skyscraper sheaf of DG-categories over ${\mathbb C}^*$ whose ``stalk'' over $h_0$ is the category of bounded complexes of vector spaces. We may denote it $D^b(Coh(\bullet))$. Including the gauge degrees of freedom means working with a sheaf of categories with a ${\mathbb C}^*$ action. Thus we simply consider a skyscraper sheaf of categories over ${\mathbb C}^*$ whose ``stalk'' over $h_0$ is the category of ${\mathbb C}^*$-equivariant complexes of vector spaces $D^b_{{\mathbb C}^*}(Coh(\bullet))$. The monoidal category $D^b_{{\mathbb C}^*}(Coh({\mathbb C}^*))$ acts on it in a fairly obvious manner: one simply tensors an object of $D^b_{{\mathbb C}^*}(Coh(\bullet))$ with the (derived) restriction of an object of $D^b_{{\mathbb C}^*}(Coh({\mathbb C}^*))$ to the point $h=h_0$. \subsection{Line operators on Gukov-Witten surface operators} The category of morphisms between two different skycraper sheaves of categories is trivial (the set of objects is empty). This corresponds to the fact that two different Gukov-Witten surface operators cannot join along a boundary-changing line operator. But the category of line operators sitting on a particular Gukov-Witten surface operator (i.e. the endomorphism category of a Gukov-Witten surface operator) is nontrivial. Its most obvious objects are Wilson lines for the complexified gauge field ${\mathcal A}$, which are obviously BRST-invariant. Such operators are labeled by irreducible representations of ${\mathbb C}^*$. One might guess therefore that the category of surface line operators is simply the category of representations of ${\mathbb C}^*$, or perhaps the category of ${\mathbb C}^*$-equivariant complexes of vector spaces which we denoted $D^b_{{\mathbb C}^*}(Coh(\bullet))$ above. However, this naive guess is wrong, which can be seen by inspecting BRST-invariant local operators which can be inserted into such a Wilson line operator. From the abstract viewpoint they form an algebra (the endomorphism algebra of an object in the category of line operators). It is clear that any power of the field $\sigma$ gives such an operator, so the algebra of local operators on a line operator is the algebra of polynomial functions of a single variable of ghost number $2$. In what follows we will denote the line parameterized by $\sigma$ by ${\mathbb C}[2]$ to indicate that $\sigma$ sits in degree $2$; thus ${\mathbb C}[2]$ is a purely even graded manifold. On the other hand, the algebra of endomorphisms of an irreducible representation of ${\mathbb C}^*$ is simply ${\mathbb C}$. To determine what the category of line operators is it is convenient to take the 2d viewpoint and reduce the 3d theory on an interval with the Gukov-Witten-type boundary condition on both ends. Let $x^3$ denote the coordinate on the interval. Gukov-Witten boundary conditions eliminate the complex scalar $h$ (which is now locked at the value $h_0$) and the field $\phi_3$ but keep the complex scalar $\sigma$ and the gauge field ${\mathcal A}$. Thus the effective 2d theory also has two sectors: the B-model with target ${\mathbb C}[2]$ and the B-type 2d topological gauge theory. According to appendix B, the corresponding category of branes is equivalent to the ${\mathbb C}^*$-equivariant derived category of ${\mathbb C}[2]$: its objects can be regarded as ${\mathbb C}^*$-equivariant complexes of holomorphic vector bundles on ${\mathbb C}[2]$ (with a trivial ${\mathbb C}^*$ action on ${\mathbb C}[2]$). This answer is independent of the parameter $h_0=\exp(-2\pi(\alpha-i\gamma))$ of the Gukov-Witten surface operator. In particular, we can choose the trivial surface operator $h_0=1$, in which case we should get the category of bulk line operators in the GL-twisted theory at $t=i$. It is not difficult to see that this answer for the category of bulk line operators agrees with the computation of the endomorphism algebra of a Wilson line explained above. Indeed, an insertion of a Wilson line does not put any constraints on $\sigma$ and does not add any degrees of freedom, and therefore should correspond to a trivial line bundle over ${\mathbb C}[2]$. Its fiber carries a representation of ${\mathbb C}^*$ determined by the charge of the Wilson line. The endomorphism algebra of such an object of $D^b_{{\mathbb C}^*}(Coh({\mathbb C}[2]))$ is simply the algebra of polynomial functions on ${\mathbb C}[2]$. It is now clear that the category of line operators contains objects other than Wilson lines. For example, we may consider a skyscraper sheaf at the origin of ${\mathbb C}[2]$, whose stalk at the origin is a complex line $V$ carrying some representation of ${\mathbb C}^*$. There are two different way to define the corresponding line operator. First, we may consider a free resolution of the skyscraper: $$ V[-2] \otimes {\mathcal O} \rightarrow V\otimes {\mathcal O}, $$ where $V[-2]$ means $V$ placed in ghost degree $-2$, ${\mathcal O}$ is the algebra of polynomial functions on ${\mathbb C}[2]$, and the cochain map is multiplication by $\sigma$. The shift by $-2$ is needed so that the cochain map has total degree $1$. The existence of such a resolution means that we can realize the ``skyscraper'' line operator as a ``bound state'' of two Wilson lines both associated with the representation $V$ but placed in different cohomological degrees. The corresponding bulk line operator is obtained using the formulas of appendix 2, where the target of the gauged B-model is taken to be ${\mathbb C}[2]$, the vector bundle $E$ on ${\mathbb C}[2]$ is trivial and of rank $2$, with graded components in degrees $1$ and $0$, and the bundle morphism $T$ from the former to the latter component is multiplication by $\sigma$. In accordance with the appendix, we consider a superconnection on $\sigma^*E$ of the form $$ {\mathcal N}=\begin{pmatrix} n{\mathcal A} & 0\\ \frac12 (\psi-i{\tilde\psi}) & n{\mathcal A}\end{pmatrix}, $$ where $n\in{\mathbb Z}$ is the weight with which ${\mathbb C}^*$ acts on $V$. The bulk line operator corresponding to the skyscraper sheaf at the origin of ${\mathbb C}[2]$ is the holonomy of this superconnection along the insertion line $\ell$. Another (equivalent) way is to take seriously the fact that the skyscraper sheaf is localized at $\sigma=0$ and require the field $\sigma$ to vanish at the insertion line $\ell$. To make this well-defined, one needs to excise a small tubular neighborhood of $\ell$ and impose a suitable boundary condition on the resulting boundary. This condition must set $\sigma=0$ and leave the components of ${\mathcal A}$ tangent to the boundary $\ell$ unconstrained. BRST-invariance determines uniquely the boundary conditions for all other fields. \section{Surface operators at $t=i$: the nonabelian case} \subsection{Reduction to 3d in the nonabelian case}\label{sec:nonabelianred} To generalize the preceding discussion to the nonabelian case we need to understand the 3d TFT which is obtained by compactifying the 4d gauge theory on a circle. This is less straightforward than in the abelian case, because requiring the fields to be independent of the $x^4$ coordinate is not a gauge-invariant condition. One can try to avoid dealing with this issue by first fixing a gauge such that $A_4$ does not depend on $x^4$. This works in the neighborhood of $A_4=0$, i.e. when the holonomy of $A$ along $S^1$ is close to $1$. But in general the condition that $A_4$ is $x^4$-independent does not fix the freedom to make $x^4$-dependent gauge transformations. For example, suppose $A_4$ is proportional to an element $\mu\in{\mathfrak g}$ which satisfies $$ \exp(2\pi \mu)=1. $$ Such $\mu$ are precisely those which lie in the $G$-orbits of the cocharacter lattice of $G$. Then the gauge transformation $$ g(x^4)=\exp(\mu x^4) $$ shifts $A_4$ by $\mu$: $$ A_4\mapsto A_4+\mu $$ Such a gauge transformation in general makes other fields $x^4$-dependent. It is shown in appendix E that the naive reduction procedure which requires all fields to be independent of $x^4$ gives the gauged Rozansky-Witten model with target $T^*[2]{\mathfrak g}_{\mathbb C}\simeq {\mathfrak g}_{\mathbb C} \times {\mathfrak g}_{\mathbb C}^*[2]$, where the gauge group $G$ acts on the base ${\mathfrak g}$ and the fiber ${\mathfrak g}^*[2]$ via the adjoint and coadjoint representations, respectively. The symplectic form is the canonical form on the cotangent bundle. The true target space of the reduced model is $T^*[2]G_{\mathbb C}$ which contains an open neighborhood of the origin in $T^*[2]{\mathfrak g}_{\mathbb C}$ as an open subset. We conjecture that the 3d theory is the gauged Rozansky-Witten model with target $T^*[2]G_{\mathbb C}$, basically because it is the only obvious possibility. \subsection{Some simple boundary conditions} Let us consider some boundary conditions in the gauged Rozansky-Witten model with target $T^*[2]G_{\mathbb C}$. The most natural boundary condition in the gauge sector is the Neumann condition, which preserves full gauge-invariance on the boundary. In the matter sector one has to pick a $G$-invariant complex Lagrangian submanifold of $T^*[2]G_{\mathbb C}$ which is invariant with respect to the rescaling of the fiber. Such a Lagrangian submanifold can be constructed by picking a $G_{\mathbb C}$-invariant closed complex submanifold of $G_{\mathbb C}$ and taking its conormal bundle. For example, one can take the whole $G_{\mathbb C}$, and then the Lagrangian submanifold is given by $\sigma=0$. We will call the resulting boundary condition in the gauged RW model the distinguished boundary condition. It is an analogue of the NN condition in the abelian case. Another natural choice of a $G$-invariant Lagrangian submanifold is the conormal bundle of a complex conjugacy class in $G_{\mathbb C}$. In order for the submanifold to be closed take the conjugacy class to be semisimple. This boundary condition is a nonabelian analogue of the ND condition. The corresponding surface operator is a semisimple Gukov-Witten-type surface operator. Indeed, fixing a semisimple conjugacy class of $\exp(-2\pi(A_4+i\phi_4))$ is the same as fixing a semisimple conjugacy class of the limiting holonomy of the complex connection $A+i\phi$ in the 4d gauge theory. More generally, if the conjugacy class is not closed, one needs to consider the conormal bundle of its closure. It is easy to analyze boundary line operators for these boundary conditions. Reducing the 3d theory on an interval with the distinguished boundary conditions we get a B-type 2d gauge theory coupled to a B-model with target $G_{\mathbb C}$. The gauge group acts on $G_{\mathbb C}$ by conjugation. According to appendix B, the corresponding category of branes is equivalent to $D^b_{G_{\mathbb C}}(Coh(G_{\mathbb C}))$. The monoidal structure cannot be deduced from the 2d considerations, but the same analysis as in the usual RW model shows that it is given by the derived tensor product. In the Gukov-Witten case we need to fix a semisimple complex conjugacy class ${\mathcal C}$ in $G_{\mathbb C}$. Let $N^*{\mathcal C}$ denote the total space of its conormal bundle in $T^*G_{\mathbb C}$. Concretely, it is the space of pairs $(g,\sigma)$, where $g\in{\mathcal C}$ and $\sigma\in{\mathfrak g}_{\mathbb C}$ satisfies ${\rm Tr}\, \sigma\, g^{-1}\delta g=0$ for any $\delta g$ tangent to ${\mathcal C}$ at $g$. The fiber coordinate $\sigma$ has cohomological degree $2$; to indicate this we will denote the corresponding graded complex manifold $N^*[2]{\mathcal C}$. Reduction on an interval in the Gukov-Witten case gives a B-type 2d gauge theory coupled to a B-model whose target is $N^*[2]{\mathcal C}$. Its category of branes is $D^b_{G_{\mathbb C}}(Coh(N^*[2]{\mathcal C}))$. The monoidal structure is given by the derived tensor product. \subsection{Bulk line operators} It is interesting to consider the special case of a Gukov-Witten surface operator corresponding to the trivial conjugacy class in $G_{\mathbb C}$ (i.e. the identity). This is the trivial surface operator, so the category of 3d boundary line operators in this case can be identified with the category of bulk line operators in the 4d TFT. The conormal bundle of the identity element is simply the dual of the complexified Lie algebra ${\mathfrak g}_{\mathbb C}$; the group $G_{\mathbb C}$ acts on it by the adjoint representation. Thus the category of 4d bulk line operators is equivalent to $D^b_{G_{\mathbb C}}(Coh({\mathfrak g}^*_{\mathbb C}[2]))$. In other words, it is the $G_{\mathbb C}$-equivariant derived category of the graded algebra $\oplus_p {\rm Sym}^p {\mathfrak g}$ where the $p^{\rm th}$ component sits in cohomological degree $2p$. In view of this result it is interesting to consider local operators sitting at the junction of two Wilson loops in representations $V_1$ and $V_2$ of $G$. The corresponding objects of the category $D^b_{G_{\mathbb C}}(Coh({\mathfrak g}^*_{\mathbb C}[2]))$ are free modules over ${\mathfrak A}=\oplus_p {\rm Sym}^p {\mathfrak g}[2]$ of the form $V_1\otimes_{\mathbb C} {\mathfrak A}$ and $V_2\otimes {\mathfrak A}$, with the obvious $G_{\mathbb C}$ action. The space of morphisms between them is the space of $G_{\mathbb C}$-invariants in the infinite-dimensional graded representation $$ V_1^*\otimes V_2\otimes {\mathfrak A} $$ Indeed, a BRST-invariant and gauge-invariant junction of two Wilson lines should be an operator in representation $V_1^*\otimes V_2$ constructed out of the complex scalar $\sigma$ taking values in ${\mathfrak g}_{\mathbb C}$. The space of such operators in ghost number $2p$ is ${\rm Hom}_G ({\rm Sym}^p {\mathfrak g},V_1^*\otimes V_2)$, where ${\rm Hom}_G$ denotes the space of morphisms in the category of representations of $G$. Summing over all $p$ we get the above answer. \subsection{More general surface operators} As in the abelian case, the above examples do not exhaust the set of objects in the 2-category of surface operators. For example, in \cite{Witten:wild} more complicated surface operators have been considered which involve higher-order poles for the complex connection ${\mathcal A}=A+i\phi$. By analogy with the Rozansky-Witten model we propose that the most general surface operator at $t=i$ (or equivalently, the most general boundary condition in the 3d theory) can be defined as a module category of the monoidal category of boundary line operators for the distinguished boundary condition ${\mathbb X}_0$. As explained above, this monoidal category is $D^b_{G_{\mathbb C}}(Coh(G_{\mathbb C}))$. Concretely, this means that the most general surface operator can be obtained by fibering a family of 2d TFTs over $G_{\mathbb C}$, so that the $G_{\mathbb C}$ action on the base (by conjugation) lifts to a $G_{\mathbb C}$ action on the whole family. For example, one may consider a complex manifolds $X$ which is a fibration over $G_{\mathbb C}$, so that fibers are Calabi-Yau manifolds, and one is given a lift of the $G_{\mathbb C}$ action on the base (by conjugation) to a $G_{\mathbb C}$ action on the total space. Given any surface operator ${\mathbb X}$, one may construct a module category over $D^b_{G_{\mathbb C}}(Coh(G_{\mathbb C}))$ by looking at the category of line operators sitting at the junction of ${\mathbb X}$ and the distinguished surface operator ${\mathbb X}_0$. This category is the category of branes in the 2d TFT obtained by compactifying the 3d TFT on an interval, with the boundary conditions on the two ends given by ${\mathbb X}$ and ${\mathbb X}_0$. Equivalently, one may compactify the 4d TFT on a twice-punctured 2-sphere, with surface operators ${\mathbb X}$ and ${\mathbb X}_0$ inserted at the two punctures. For example, if we consider a surface operator defined, as in \cite{Witten:wild}, by a prescribed singularity in the complex connection ${\mathcal A}$, and take into account that the distinguished surface operator is defined by allowing the holonomy of ${\mathcal A}$ to be free, we see that the space of vacua of the effective 2d TFT is the moduli space of connections on a punctured disc with the prescribed singularity at the origin. Let us denote this moduli space ${\mathcal M}$. If in the definition of ${\mathcal M}$ we divide by the group of gauge transformations which reduce to the identity at some chosen point on the boundary of the disk, then ${\mathcal M}$ is acted upon by $G_{\mathbb C}$ and is fibered over $G_{\mathbb C}$ (the holonomy of ${\mathcal A}$ along the boundary of the disk). It looks plausible that the effective 2d TFT is the B-model with target ${\mathcal M}$ coupled to a B-type gauge theory with gauge group $G_{\mathbb C}$. Its category of branes is a module category over $D^b_{G_{\mathbb C}}(Coh(G_{\mathbb C}))$. \section{Surface operators at $t=1$} \subsection{Reduction to 3d} For $t=1$ the 4d TFT compactified on a circle also decomposes into two sectors (gauge and matter), but the analysis of boundary conditions is less straightforward because neither sector has been studied previously. For this reason we will restrict ourselves to the abelian case, which is fairly elementary. The gauge sector consists of a 3d gauge field $A$, a real bosonic 0-form $\phi_4$, a complex bosonic 0-form $\sigma$, a fermionic 2-form $\chi=\frac12 \chi_{ij} dx^i dx^j$, a fermionic 1-form $\psi+{\tilde\psi},$, and fermionic 0-forms $\psi_4-{\tilde\psi}_4,\eta+{\tilde\eta}$. Their BRST transformations read \begin{align*} \delta A &=i(\psi+{\tilde\psi}),\\ \delta\phi_4 &=i(\psi_4-{\tilde\psi}_4),\\ \delta\sigma &=0,\\ \delta{\bar\sigma} &= i(\eta+{\tilde\eta}),\\ \delta (\psi+{\tilde\psi}) &=2d\sigma,\\ \delta(\psi_4-{\tilde\psi}_4) &=0,\\ \delta(\eta+{\tilde\eta}) &=0,\\ \delta\chi &=F+\star d\phi_4 \end{align*} On-shell they satisfy $\delta^2=2i\delta_g(\sigma)$, where $\delta_g(\sigma)$ is a gauge transformation with the parameter $\sigma$. The action is BRST-exact: $$ S_{gauge}=-\frac{1}{2e^2}\,\delta \int_{M_3} \left(\chi\wedge\star (F+\star d\phi_4)+\frac12 (\psi+{\tilde\psi})\wedge \star d{\bar\sigma}\right). $$ The matter sector consists of a real periodic scalar $A_4$, a real bosonic 1-form $\phi$, fermionic 1-forms $\psi-{\tilde\psi},\rho=\chi_{i4}dx^i$, and fermionic 0-forms $\psi_4+{\tilde\psi}_4,\eta-{\tilde\eta}$. Their BRST transformations read \begin{align*} \delta A_4 &=i(\psi_4+{\tilde\psi}_4),\\ \delta \phi &=i(\psi-{\tilde\psi}),\\ \delta(\psi-{\tilde\psi}) &=0,\\ \delta(\psi_4+{\tilde\psi}_4)&=0,\\ \delta\rho &=dA_4+\star d\phi,\\ \delta(\eta-{\tilde\eta})&=2 d^\star\phi. \end{align*} On-shell they satisfy $\delta^2=0$. The matter action is also BRST-exact: $$ S_{matter}=-\frac{1}{2e^2}\, \delta \int_{M_3} \left(\rho\wedge\star (dA_4+\star d\phi)+\frac12 (\eta-{\tilde\eta})\wedge \star d^\star\phi\right). $$ \subsection{Boundary conditions in the gauge sector} The gauge sector is the dimensional reduction of the Donaldson-Witten 4d TFT \cite{Witten:Donaldson} down to 3d. Above we have called this theory an A-type gauge theory. However, this by itself does not teach us very much, since boundary conditions in this theory have not been discussed previously. Without adding boundary degrees of freedom, the only choices are the Dirichlet and Neumann boundary conditions for gauge fields, with BRST-invariance fixing the conditions on all other fields. \subsubsection{The Dirichlet condition} Let us begin with the Dirichlet condition which says that the restriction of $A$ to the boundary is trivial. Since $\delta^2=2i\delta_g(\sigma)$, this makes sense only if $\sigma$ also vanishes on the boundary. BRST-invariance then requires $\eta+{\tilde\eta}$ and the restriction of the 1-form $\psi+{\tilde\psi}$ to vanish. The fermionic equations of motion then require the restriction of $\chi$ to vanish, and the BRST-invariance implies that $\phi_4$ must satisfy the Neumann condition $\partial_3\phi_4=0$, where we assumed that the boundary is given by $x^3=0$. The Dirichlet boundary condition has the property that it has no nontrivial local BRST-invariant boundary observables. Indeed, the only nonvanishing BRST-invariant 0-form is $\psi_4-{\tilde\psi}_4$, but it is BRST-exact. To analyze boundary line operators, we use the dimensional reduction trick and compactify the 3d theory on an interval with the Dirichlet boundary conditions. The only bosonic fields in the effective 2d theory are the constant mode of $\phi_4$ and the holonomy of $A$ along the interval parameterized by $x^3$. That is, the bosonic fields are a real scalar and a periodic real scalar. The effective 2d TFT is therefore a sigma-model with target ${\mathbb R}\times S^1$. In fact, it can be regarded as an A-model with target $T^*S^1$. The easiest way to see this is to note that the path-integral of the 3d theory localizes on configurations given by solutions of the Bogomolny equations $$ F+\star d\phi_4=0. $$ Upon setting all fields to zero except $A_3$ and $\phi_4$ and assuming that they are independent of $x^3$, this equation becomes $$ dA_3+\medstar d\phi_4=0, $$ where $\medstar$ is the 2d the Hodge star operator. This is an elliptic equation which can be interpreted as the holomorphic instanton equation, provided we declare $A_3+i\phi_4$ to be a complex coordinate on the target. Since the action of the 4d theory is BRST-exact, so is the action of the 2d model. This agrees with the well-known fact that the action of an A-model is BRST-exact if the symplectic form on the target space is exact. The category of line operators on the Dirichlet boundary is therefore the Fukaya-Floer category of $T^*S^1$ whose simplest objects are Lagrangian submanifolds equipped with unitary vector bundles with flat connections. Since this category arises as the endomorphism category of an object in a 2-category, it must have a monoidal structure, which is not visible from the purely 2d viewpoint. In fact, we do not expect the Fukaya-Floer category of a general symplectic manifold to have a natural monoidal structure. We will argue below that the monoidal structure is induced by the mirror symmetry which establishes the equivalence of the Fukaya-Floer category of $T^*S^1$ with $D^b(Coh({\mathbb C}^*))$ and the monoidal structure on the latter category. For now we just note that the base $S^1$ has a distinguished point corresponding to the trivial holonomy of $A$ on the interval. The fiber over this point is a Lagrangian submanifold in $T^*S^1$ and is the identity object with respect to the monoidal structure. The distinguished point allows us to identity $S^1$ with the group manifold $U(1)$. \subsubsection{The Neumann condition} Now let us consider the Neumann condition for the 3d gauge field $A$. This means that the gauge symmetry is unbroken on the boundary and the restriction of the 1-form $\star F$ vanishes. Then the Bogomolny equation requires $\phi_4$ to have the Dirichlet boundary condition $\phi_4=a=const$, and by BRST-invariance $\psi_4-{\tilde\psi}_4$ must vanish at $x^3=0$. Fermionic equations of motion imply then that $\psi_3+{\tilde\psi}_3$ vanishes as well, and since $\delta(\psi_3+{\tilde\psi}_3)=2\partial_3\sigma$, the field $\sigma$ satisfies the Neumann condition. Finally, the restriction of the 1-form $\star \chi$ to the boundary must vanish, in order for the fermionic boundary conditions to be consistent. Indeed, if $x^1$ is regarded as the time direction, then $(\star \chi)_2$ is canonically conjugate to $\psi_3+{\tilde\psi}_3$, so if one of them vanishes, so should the other. Similarly, if $x^2$ is regarded as time, then $(\star\chi)_1$ is canonically conjugate to $\psi_3+{\tilde\psi}_3$ and therefore must vanish too. In the Neumann case the space of BRST-invariant local observables on the boundary is spanned by powers of the field $\sigma$. To determine the category of boundary line operators one has to reduce the 3d gauge theory on an interval with the Neumann boundary conditions. The bosonic fields of the effective 2d theory are the 2d gauge field and the constant mode of the scalar $\sigma$, the fermionic ones are the 0-form $\eta+{\tilde\eta}$, the 1-form $\psi+{\tilde\psi}$, and the 2-form $\chi$. Their BRST transformations are \begin{align*} \delta A &=i(\psi+{\tilde\psi}),\\ \delta \sigma &=0,\\ \delta{\bar\sigma} &=i(\eta+{\tilde\eta}),\\ \delta(\eta+{\tilde\eta}) &=0,\\ \delta (\psi+{\tilde\psi}) & =2d\sigma,\\ \delta \chi &= F. \end{align*} This 2d TFT can be obtained from the usual $N=(2,2)$ $d=2$ supersymmetric gauge theory by means of a twist which makes use of the $U(1)_V$ R-symmetry. Since this is the same R-symmetry as that used for constructing an A-type sigma-model, we might call this TFT an A-type 2d gauge theory. As far as we know, its boundary conditions have not been analyzed in the literature previously. It is shown in appendix F that its category of branes is equivalent to the bounded derived category of coherent sheaves on the graded line ${\mathbb C}[2]$. \footnote{This category is equivalent to the $U(1)$-equivariant constructible derived category of sheaves over a point \cite{BL}.} Again, the 3d origin of this category means that it must have monoidal structure. Here it is given by the usual derived tensor product of complexes of coherent sheaves. The trivial line bundle on ${\mathbb C}[2]$ is the identity object. From the 3d viewpoint, it corresponds to the ``invisible" line operator on the boundary. As mentioned above, the Neumann condition depends on a real parameter $a$, the boundary value of the scalar $\phi_4$. On the quantum level there is another parameter which takes values in ${\mathbb R}/2\pi {\mathbb Z}$. It enters as the coefficient of a topological term in the boundary action: $$ \theta\int_{x^3=0} \frac{F}{2\pi}. $$ Thus overall the Neumann condition in the gauge sector has the parameter space ${\mathbb R}\times\SS^1\simeq{\mathbb C}^*$. \subsection{Boundary conditions in the matter sector} We may impose either Dirichlet or Neumann condition on the periodic scalar $A_4$. Let us discuss these two possibilities in turn. \subsubsection{The Dirichlet condition} If $A_4$ satisfies the Dirichlet condition, then BRST-invariance requires the 1-form $\phi$ to satisfy the Neumann condition. This means that the components of $\phi$ tangent to the boundary are free and satisfy $\partial_3\phi_1=\partial_3\phi_2=0$, while the component $\phi_3$ takes a fixed value $\phi_3=a$ on the boundary. BRST-invariance also requires the following fermions to vanish on the boundary: $\psi_4+{\tilde\psi}_4$, $\psi_3-{\tilde\psi}_3$, $\rho_1,$ $\rho_2$. The real parameter $a$ together with the boundary value of $A_4$ combine into a parameter taking values in $S^1\times{\mathbb R}$. These parameters are actually irrelevant, in the sense that topological correlators do not depend on them. To see this, note that shifting the boundary value of $A_4$ can be achieved by adding a boundary term to the action of the form $$ \int_{x^3=0} \partial_3 A_4 d^2x =\int_{x^3=0} \left(\delta\rho_3-(\partial_1\phi_2-\partial_2\phi_1)\right) d^2x $$ We see that up to a total derivative this boundary terms is BRST-exact, hence does not affect the correlators. A similar argument can be made for the boundary value of $\phi_3$. The reduction on an interval with the Dirichlet boundary conditions gives rise to a 2d TFT whose only bosonic field is a real 1-form $\phi$. Such a 2d TFT has not been considered previously, but it is closely related to an A-model with target $T^*{\mathbb R}$. To see this, consider an $N=(2,2)$ supersymmetric sigma-model with target ${\mathbb C}$ (with the standard flat metric). This model has a $U(1)$ symmetry which acts on the target space coordinate $Z$ by $$ Z\mapsto e^{i\alpha} Z. $$ One can add a multiple of the corresponding $U(1)$ current to the standard R-current, thereby defining a new R-current. When performing the A-twist, we can choose this modified R-current instead of the standard one. If $Z$ has charge two with respect to the modified R-symmetry, after twist ${\rm Re}\ Z$ and ${\rm Im}\ Z$ will become components of a 1-form. We will call the resulting 2d TFT the modified A-model. Apart from the bosonic 1-form $\phi$, the modified A-model has a fermionic 1-form $\psi-{\tilde\psi}$ and a pair of fermionic 0-forms $\eta-{\tilde\eta}$ and $\rho$ (the latter comes from the component $\rho_3$ of the 1-form $\rho$ in 3d). Their BRST transformations are \begin{align*} \delta\phi &=i(\psi-{\tilde\psi}),\\ \delta (\psi-{\tilde\psi}) &=0,\\ \delta(\eta-{\tilde\eta}) &=2d^\medstar\phi,\\ \delta\rho &=\medstar d\phi \end{align*} Here $\medstar$ is the 2d Hodge star operator, and $d^\medstar=\medstar d\medstar$. To understand the category of boundary line operators in 3d, we need to describe the category of boundary conditions for the modified A-model. This is fairly straightforward. A natural class of boundary conditions is obtained by imposing on the boundary $$ \left(a \phi+b \medstar \phi\right)\vert_{\partial M_2}=0. $$ The special cases $b=0$ and $a=0$ correspond to the 2d Dirichlet and Neumann conditions. Since the theory obviously has a symmetry rotating $\phi$ into $\medstar\phi$, it is sufficient to consider the Neumann condition $\medstar\phi\vert=0$. BRST-invariance requires the restriction of $\medstar\psi$ and $\rho$ to vanish on such a boundary. It is easy to see that there are no nontrivial BRST-invariant boundary observables (the only BRST-invariant fermion $\psi$ is BRST-exact), so there is no possibility to couple boundary degrees of freedom in a nontrivial way. This implies that the category of boundary conditions is the same as for a trivial 2d TFT, i.e. the category of complexes of finite-dimensional vector spaces. We may denote it $D^b(Coh(\bullet))$. There is an important subtlety here related to the fact that the scalar $A_4$ is periodic with period $1$. When reducing on an interval, this means that there are ``winding sectors'', where $$ \int dx^3 \partial_3 A_4= n,\quad n\in{\mathbb Z}. $$ This winding is constant along a connected component of the boundary and does not affect the 2d theory in any way. We may incorporate it by introducing an additional integer label on each boundary component which serves as a conserved boundary charge. This is mathematically equivalent to saying that the category of boundary conditions is the category of ${\mathbb C}^*$-equivariant coherent sheaves over a point $D^b_{{\mathbb C}^*}(Coh(\bullet))$. Objects of this category are complexes of finite-dimensional vector spaces with a ${\mathbb C}^*$-action, such that the differentials in the complex commute with the ${\mathbb C}^*$ action. Morphisms are required to preserve the ${\mathbb C}^*$-action, i.e. to have zero ${\mathbb C}^*$-charge. \subsubsection{The Neumann condition} If $A_4$ satisfies the Neumann condition $\partial_3A_4=0$, then BRST-invariance requires $\phi$ to satisfy the Dirichlet condition . That is, the restriction of $\phi$ to the boundary must vanish, and $\phi_3$ must satisfy $\partial_3\phi_3=0$. This boundary condition does not have any parameters. The reduction on an interval gives rise to the A-model with the bosonic fields $A_4$ and $\phi_3$. This can be seen for example by looking at the 3d BPS equation $dA_4+\star d\phi=0$ and restricting to field configurations where $\phi_1=\phi_2=0$ and $A_4$ and $\phi_3$ are independent of $x^3$. For such field configuration the BPS equation becomes the holomorphic instanton equation with target $S^1\times{\mathbb R}\simeq {\mathbb C}^*$. From the symplectic viewpoint, ${\mathbb C}^*$ with its standard K\"ahler form is isomorphic to $T^*S^1$. Thus the category of boundary line operators in this case is the Fukaya-Floer category of $T^*S^1$. Since this category arises as the category of boundary line operators in the 3d TFT, it must have a monoidal structure. Although the category appears to be the same as in the gauge sector with the Dirichlet boundary condition, we will see that the monoidal structure is completely different and is induced by the equivalence between (a version of) the Fukaya-Floer category of $T^*S^1$ and the constructible derived category of $S^1$ \cite{ZN}. In particular, the identity object (i.e. the invisible boundary line operator) is different and corresponds to the zero section of $T^*S^1$ with a trivial rank-1 local system. This illustrates the fact that a monoidal structure on branes in a 2d TFT depends on the way this 2d TFT is realized as a compactification of a 3d TFT on an interval. \subsection{Electric-magnetic duality} We are now ready to describe how the 4d electric-magnetic duality acts on various boundary conditions described above. Since for both gauge and matter sectors one can have either Dirichlet or Neumann conditions, there are four possibilities to consider. From the 3d viewpoint, 4d electric-magnetic duality amounts to dualizing the 3d gauge field $A$ into a periodic scalar, and simultaneously dualizing the periodic scalar $A_4$ into a 3d gauge field. It is easy to see that electric-magnetic duality applied to the A-type gauge theory gives the Rozansky-Witten model with target $T^*[2]{\mathbb C}^*$, i.e. it maps the A-type gauge sector to the B-type matter sector. Similarly, it maps the A-type matter sector into the B-type gauge theory (with gauge group $U(1)$). In other words, electric-magnetic duality reduces to particle-vortex duality done twice. The dual of the Neumann condition for a periodic scalar is the Dirichlet condition for the gauge field, and vice-versa. We will use this well-known fact repeatedly in what follows. \subsubsection{The DD condition} The first possibility is the Dirichlet condition in both gauge and matter sectors at $t=1$. The Dirichlet condition in the A-type gauge sector maps into a boundary condition in the Rozansky-Witten model with target $T^*[2]{\mathbb C}^*$ which sets $\sigma=0$ on the boundary and leaves the complex scalar $\tau$ free to fluctuate. The Dirichlet condition in the A-type matter sector is mapped to the Neumann condition in the B-type gauge theory. Note that the Dirichlet condition in the A-type matter sector has two real parameters taking values in $S^1$ and ${\mathbb R}$. The former one is mapped to a boundary theta-angle, i.e. a boundary term in the action of the form $$ \theta\int_{x^3=0} \frac{F}{2\pi}=\theta \int_{x^3=0} \frac{{\mathcal F}}{2\pi}. $$ The latter parameter is the boundary value of the field $\phi_3$. Both of these parameters are irrelevant, as discussed in section 4. As discussed above, the category of boundary line operators in the A-type 3d gauge theory is the Fukaya-Floer category of $T^*U(1)$. On the other hand, the category of boundary line operators in the Rozansky-Witten model is $D^b(Coh({\mathbb C}^*))$, as explained in \cite{KRS}. These categories are equivalent, by the usual 2d mirror symmetry. Let us recall how 2d mirror symmetry acts on some objects in this case. The trivial line bundle on ${\mathbb C}^*$ is mapped to the fiber over a distinguished point of the base $S^1$. This distinguished point allows us to identify $S^1$ with the group manifold $U(1)$. More generally, we may consider a holomorphic line bundle on ${\mathbb C}^*$ with a ${\bar\partial}$-connection of the form $$ {\bar\partial}+i\lambda\frac{d{\bar z}}{{\bar z}},\quad \lambda\in{\mathbb C}. $$ We will denote such a line bundle ${\mathcal L}_\lambda$. Gauge transformations can be used to eliminate the imaginary part of $\lambda$. They also can shift the real part of $\lambda$ by an arbitrary integer. Thus we may regard the parameter $\lambda$ as taking values in ${\mathbb R}/{\mathbb Z}\simeq S^1$. Mirror symmetry maps ${\mathcal L}_\lambda$ to a Lagrangian submanifold in $T^*U(1)$ which is a fiber over the point $\exp(2\pi i\lambda)\in U(1)$. Applying mirror symmetry to the obvious monoidal structure on $D^b(Coh({\mathbb C}^*))$ given by the derived tensor product we get a monoidal structure on the Fukaya category of $T^*U(1)$. The trivial holomorphic line bundle on ${\mathbb C}^*$, which serves as the identity object in $D^b(Coh({\mathbb C}^*))$, is mapped to the Lagrangian fiber over the identity element of $U(1)$. If we consider two Lagrangian fibers over the points $\exp(2\pi i \lambda_1), \exp(2\pi i\lambda_2)\in U(1)$, their mirrors are line bundles ${\mathcal L}_{\lambda_1}$ and ${\mathcal L}_{\lambda_2}$. Their tensor product is a line bundle ${\mathcal L}_{\lambda_1+\lambda_2}$ whose mirror is the Lagrangian fiber over the point $\exp(2\pi i(\lambda_1+\lambda_2))\in U(1)$. Clearly, this rule for tensoring objects of the Fukaya category makes use of the group structure of $U(1)$, i.e. it is a convolution-type tensor product. Another natural class of Lagrangian submanifolds to consider are constant sections of $T^*U(1)$, i.e. submanifolds given by the equation $\phi_4=const$. These submanifolds are circles and may carry a nontrivial flat connection. Thus such A-branes are labeled by points of ${\mathbb R}\times U(1)\simeq {\mathbb C}^*$. The mirror objects are skyscraper sheaves on ${\mathbb C}^*$. The derived tensor product of two skyscrapers supported at different points is obviously the zero object. The derived tensor product of a skyscraper with itself can be shown to be isomorphic to the sum of the skyscraper and the skyscraper shifted by $-1$. That is, it is a skyscraper sheaf over the same point whose stalk is a graded vector space ${\mathbb C}[-1]\oplus {\mathbb C}$. Applying mirror symmetry, we see that the tensor product of a section of $T^*U(1)$ with itself must be the sum of two copies of the same section, but with the Maslov grading of one of them shifted by $-1$. We do not know how to reproduce this result without appealing to mirror symmetry, i.e. by computing the product of boundary line operators in the A-type gauge theory. As discussed above, the category of boundary line operators in the A-type matter sector is the category of branes in a somewhat unusual 2d TFT which is a modification of the A-model with target $T^*{\mathbb R}$. It was argued above that this category is equivalent to $D^b_{{\mathbb C}^*}(Coh(\bullet))$. This agrees with the B-side, where the reduction on an interval gives a B-type 2d gauge theory. Putting the gauge and matter sectors together, we see that the DD boundary condition on the A-side is mapped to what we called the distinguished boundary condition on the B-side. The category of boundary line operators for such a boundary condition is the ${\mathbb C}^*$-equivariant derived category of coherent sheaves $D^b_{{\mathbb C}^*}(Coh({\mathbb C}^*))$ with its obvious monoidal structure. On the A-side we get a graded version of the Fukaya-Floer category of $T^*U(1)$ where a flat vector bundle over a Lagrangian submanifold has an additional integer grading and morphisms are required to have degree zero with respect to it. This grading arises from the winding number of the periodic scalar $A_4$. We can also interpret the duality in 4d terms. Indeed, it is easy to see that the DD boundary condition on the A-side arises from a 4d Dirichlet boundary condition at $t=1$, while its dual on the B-side arises from the 4d Neumann condition at $t=i$. Thus electric-magnetic duality exchanges Dirichlet and Neumann boundary conditions in 4d, as expected. The surface operators corresponding to such 4d boundary conditions can be interpreted as follows: we excise a tubular neighborhood of the support of the surface operator and impose the 4d boundary condition on the resulting boundary. In a TFT, such a procedure gives a surface operator (i.e. there is no need to take the limit where the thickness of the tubular neighborhood goes to zero). \subsubsection{The NN condition} This condition is the distinguished boundary condition on the A-side, since the gauge group is unbroken on the boundary, and the periodic scalar $A_4$ is free to explore the whole circle. It is mapped by electric-magnetic duality to the Dirichlet boundary condition for the B-type gauge theory and the boundary condition in the RW model with target ${\mathbb C}^*$ which fixes the ${\mathbb C}^*$-valued scalar $\tau$ and leaves $\sigma$ free. Note that both the Neumann boundary condition in the A-type gauge theory and the corresponding boundary condition in the RW model have a parameter taking values in ${\mathbb C}^*\simeq{\mathbb R}\times U(1)$. Let us compare the categories of boundary line operators. The category of boundary line operators in the A-type gauge theory is the bounded derived category of coherent sheaves $D^b(Coh({\mathbb C}[2]))$. The category of boundary line operators in the RW model is also $D^b(Coh({\mathbb C}[2]))$. The category of boundary line operators in the A-type matter sector is the Fukaya-Floer category of $T^*S^1$. The category of boundary line operators in the B-type gauge sector is $D^b(Coh({\mathbb C}^*))$. Their equivalence is a special case of the usual 2d mirror symmetry. But there is more: we expect that the categories of boundary line operators are equivalent as monoidal categories. This is easy to see directly for the RW model with target ${\mathbb C}^*$ and A-type gauge theory with gauge group $U(1)$. Indeed, in both cases typical objects in the category of boundary line operators are complexes of holomorphic vector bundles which can be represented by Wilson line operators on the boundary for some superconnection on the pull-back vector bundle. In the classical approximation, fusing two such boundary line operators corresponds to the tensor product of complexes, and there can be no quantum corrections to this result. It is more complicated to compare the monoidal structures for the other pair of dual theories (B-type gauge theory and A-type matter). We will not attempt to do an independent computation on the A-side but instead describe the monoidal structure on the B-side and then explain what it corresponds to on the A-side. Note that since ${\mathbb C}^*$ is a complex Lie group, the category $D^b(Coh({\mathbb C}^*))$ has two natural monoidal structures: the derived tensor product, and the convolution-type product. The former one does not make use of the group structure, while the latter one does. The identity object of the former one is the sheaf of holomorphic functions on ${\mathbb C}^*$, while for the latter structure it is the skyscraper sheaf at the identity point $1\in{\mathbb C}^*$. It is the latter monoidal structure which describes the fusion of boundary line operators on the B-side. Indeed, the 3d meaning of the coordinate on ${\mathbb C}^*$ is the holonomy of the connection $A+i\phi$ along a small semi-circle with both ends on the boundary and centered at the boundary line operator (see figure 8). \begin{figure}[htbp] \label{fig:SOGL_boundaryline} \centering \includegraphics[height=2in]{boundaryline} \caption{A skyscraper sheaf corresponds to a boundary line operator for which the holonomy of $A+i\phi$ along a small semi-circle around it is fixed. The dot marks the location of the boundary line operator, which we view here in cross-section.} \end{figure} Skyscraper sheaves correspond to boundary line operators for which this holonomy is fixed. In particular, the skyscraper sheaf at $1\in{\mathbb C}^*$ corresponds to the ``invisible'' boundary line operator for which this holonomy is trivial. By definition, this is the identity object in the monoidal category of boundary line operators. Mirror symmetry maps a skyscraper sheaf on ${\mathbb C}^*$ to a Lagrangian submanifold of $T^*S^1$ which is a graph of a closed 1-form $\alpha$ on $S^1$. Topologically this submanifold is a circle and is equipped with a trivial line bundle with a flat unitary connection. The moduli space of such an object is ${\mathbb C}^*$: for $\lambda\in{\mathbb C}^*$ the phase of $\lambda$ determines the holonomy of the unitary connection, while the absolute value determines the integral of $\alpha$ on $S^1$. Thus the identity object on the B-side is mirror to the zero section of $T^*S^1$ with a trivial flat connection. To describe the monoidal structure on the A-side it is best to recall a theorem of Nadler \cite{Nadler} according to which (a version of) the Fukaya-Floer category of $T^*X$ is equivalent to the constructible derived category of $X$. Recall that a constructible sheaf on a real manifold $X$ is a sheaf which is locally constant on the strata of a Whitney stratification of $X$; such sheaves can be regarded as generalizations of flat connections. Objects of the constructible derived category are bounded complexes of sheaves whose cohomology sheaves are constructible. The constructible derived category has an obvious monoidal structure arising from the tensor product of complexes of sheaves. The sheaf of locally-constant functions is the identity object with respect to this monoidal structure. According to \cite{ZN,Nadler}, this object corresponds to the zero section of $T^*S^1$ with a trivial flat connection. This suggests that the monoidal structure on the A-side is given by the tensor product on the constructible derived category. It is easy to check that this is compatible with the way mirror symmetry acts on the skyscraper sheaves on ${\mathbb C}^*$. We can try put the gauge and matter sectors together. On the B-side, we have the B-model with target ${\mathbb C}^*\times {\mathbb C}[2]$ whose category of branes is $D^b(Coh({\mathbb C}^*\times{\mathbb C}[2]))$. On the A-side, we have an A-model with target $T^*S^1$ tensored with an A-type 2d gauge theory with gauge group $U(1)$. One could guess that the corresponding category of branes is a $U(1)$-equivariant version of the Fukaya-Floer category of $T^*S^1$. More generally, one could guess that the category of branes in an A-model with target $T^*X$ tensored with the A-type 2d $U(1)$ gauge theory is a $U(1)$-equivariant version of the Fukaya-Floer category of $T^*X$. It is not clear to us how to define such an equivariant Fukaya-Floer category mathematically. Given the results of \cite{ZN,Nadler}, a natural guess is the equivariant constructible derived category of sheaves on $X$. As a check, note that when $X$ is a point, the $U(1)$-equivariant constructible derived category is equivalent to $D^b(Coh({\mathbb C}[2]))$ \cite{BL}. As mentioned above and explained in appendix F, this is indeed the category of branes for the A-type 2d gauge theory. The monoidal structure seems to be the standard one (derived tensor product). On the B-side, on the other hand, the monoidal structure is a combination of the tensor product of coherent sheaves on ${\mathbb C}[2]$ and the convolution product on ${\mathbb C}^*$. \subsubsection{The DN condition} Next consider the boundary condition on the A-side which is a combination of the Dirichlet condition in the gauge sector and the Neumann condition for $A_4$ in the matter sector. It is dual to the Dirichlet condition for the B-type gauge sector and a boundary condition for the RW model with target $T^*[2]{\mathbb C}^*$ which sets $\sigma=0$ and leaves the complex scalar $\tau=A_4+i\phi_4$ free to fluctuate. On the B-side reduction on an interval gives a B-model with target ${\mathbb C}^*\times{\mathbb C}^*$, therefore the category of boundary line operators is $D^b(Coh({\mathbb C}^*\times{\mathbb C}^*))$. On the A-side reduction gives an A-model with target $T^*U(1)\times T^*U(1)$, therefore the category of boundary line operators is the Fukaya-Floer category. The two categories are equivalent by the usual 2d mirror symmetry. The monoidal structure is easiest to determine on the B-side. It is neither the derived tensor product, nor the convolution, but a combination of both. This happens because the two copies of ${\mathbb C}^*$ have a very different origin: one of them arises from a 3d B-type gauge theory, and the other one arises from the Rozansky-Witten model with target $T^*[2]{\mathbb C}^*$. \subsubsection{The ND condition} Finally we consider the boundary condition on the A-side which is a combination of the Neumann condition in the gauge sector and the Dirichlet condition for $A_4$. This is the case which corresponds to the Gukov-Witten surface operator at $t=1$. Indeed, the Dirichlet conditions for $A_4,\phi_4$ and $\phi_3$ mean that the holonomy of $A$ is fixed, while the 1-form $\phi$ has a singularity of the form $$ \beta \frac{dr}{r}-\gamma d\theta, $$ where $-\gamma$ is the boundary value of $\phi_4$ and $\beta$ is the boundary value of $\phi_3$. The boundary value of $A_4$ is the Gukov-Witten parameter $\alpha$. The Neumann condition in the gauge sector also depends on the boundary theta-angle which corresponds to the Gukov-Witten parameter $\eta$. As explained above, the boundary values of $A_4$ and $\phi_3$ are actually irrelevant. This agrees with the results of \cite{GW}, where it is shown that at $t=1$ the parameters $\alpha$ and $\beta$ are irrelevant. Thus the true parameter space of the surface operator on the A-side is ${\mathbb C}^*$. Electric-magnetic duality maps the DD condition to the Neumann condition for the B-type gauge theory and the boundary condition in the RW model which fixes $\tau$ and leaves $\sigma$ free to fluctuate. The latter boundary condition depends on the boundary value of the field $\tau=A_4+i\phi_4$. From the 4d viewpoint this boundary value encodes the Gukov-Witten parameters $\alpha$ and $\gamma$. These are the relevant parameters at $t=i$, as explained in \cite{GW}. The Neumann boundary condition in the B-type gauge theory also has two parameters (the boundary value of $\phi_3$ and the boundary theta-angle) which correspond to the Gukov-Witten parameters $\beta$ and $\eta$. But as explained above and from a different viewpoint in \cite{GW}, these parameters are irrelevant at $t=i$. Let us compare the categories of 3d boundary line operators, which from the 4d viewpoint are interpreted as categories of line operators sitting on Gukov-Witten surface operators. On the B-side reduction on an interval gives a B-model with target ${\mathbb C}[2]$ tensored with a B-type 2d gauge theory, therefore the category of boundary line operators is $D^b_{{\mathbb C}^*}(Coh({\mathbb C}[2]))$. On the A-side reduction on an interval gives an A-type 2d gauge theory tensored with a modified A-model with target $T^*{\mathbb R}$. Its category of branes is a modification of the category of boundary conditions for the A-type 2d gauge theory where the space of boundary degrees of freedom has additional integer grading coming from the winding of the periodic scalar $A_4$, and morphisms are required to have degree zero with respect to it. Since branes in the A-type 2d gauge theory can be identified with objects of $D^b(Coh({\mathbb C}[2]))$, the category of boundary conditions in the combined system is equivalent to $D^b_{{\mathbb C}^*}(Coh({\mathbb C}[2]))$, in agreement with what we got on the B-side. \subsection{A proposal for the 2-category of surface operators at $t=1$} By analogy with the Rozansky-Witten model, one may conjecture that the 2-category of surface operators at $t=0$ can be described in terms of module categories over the monoidal category of boundary line operators for the distinguished boundary condition (the NN condition). We have argued above that this monoidal category is the $U(1)$-equivariant constructible derived category of $S^1$, where the $U(1)$ action on $S^1$ is trivial. It is probably better to think about it as a sheaf of $U(1)$-equivariant monoidal DG-categories over $S^1$. To each surface operator we may associate a sheaf of $U(1)$-equivariant module categories over this sheaf of $U(1)$-equivariant monoidal categories, and we conjecture that this map is an equivalence of 2-categories. Gukov-Witten-type operators correspond to skyscraper sheaves on $S^1$. Electric-magnetic duality then implies that there is an equivalence between this 2-category and the 2-category of coherent ${\mathbb C}^*$-equivariant derived categorical sheaves over ${\mathbb C}^*$. \section{Surface operators at $t=0$} \subsection{Reduction to 3d} The 3d theory again decomposes into the gauge and matter sectors. Let us start with the gauge sector. The bosonic fields are a gauge field $A$, a periodic scalar $A_4$, and a complex scalar $\sigma$. The fermionic fields are two 0-forms $\eta$ and $\psi_4$, a 1-form $\psi$ and a 2-form $\chi_+$. Thus subscript $+$ indicates that $\chi_+$ originates from the self-dual part of the 2-form $\chi$ in four dimensions.The BRST transformations are \begin{align*} \delta A &=i\psi,\\ \delta A_4 &=i\psi_4,\\ \delta\psi &=d\sigma,\\ \delta\psi_4 &=0,\\ \delta\sigma &=0,\\ \delta{\bar\sigma} &=i\eta,\\ \delta\eta &=0,\\ \delta\chi_+ &= F+\star dA_4 \end{align*} The field content and BRST transformations are the same as in the A-type 3d gauge theory, the main difference being that the bosonic scalar $A_4$ is periodic. The action of the gauge sector contains, apart from a BRST-exact term, a topological term \begin{equation}\label{topaction} S_{top}=-\frac{2\pi}{e^2}\int_{M_3} F\wedge dA_4 \end{equation} Note that it is the periodicity of $A_4$ that makes this topological term nontrivial in general. The above topological term term comes from the dimensional reduction of a topological term in 4d $$ -\frac{1}{2e^2}\int F\wedge F. $$ Here we assumed that the 4d theta-angle vanishes and that the coordinate $x^4$ has period $2\pi$. In the matter sector the only bosonic fields are a 0-form $\phi_4$ and a 1-form $\phi$. The fermionic fields are a pair of 0-forms ${\tilde\eta}$ and ${\tilde\psi}_4$, a 1-form ${\tilde\psi}$, and a 2-form $\chi_-$ which arises from the anti-self-dual part of the 2-form $\chi$ in four dimensions. The matter content and BRST transformations are the same as for the $t=1$ matter sector, except that the periodic scalar $A_4$ is replaced with a non-periodic scalar $\phi_4$. The matter action is BRST-exact. \subsection{The gauge sector} As for $t=1$, we may consider either Dirichlet or Neumann conditions for the gauge field, and then BRST-invariance determines the rest. The category of boundary line operators is determined by compactifying the theory on an interval with the appropriate boundary conditions and analyzing branes in the resulting 2d TFT. In the Neumann case the effective 2d TFT is the A-type 2d gauge theory, just as for $t=1$. As explained above, its category of boundary conditions is equivalent to $D^b(Coh({\mathbb C}[2]))$. In the Dirichlet case the effective 2d TFT is a topological sigma-model with two bosonic fields, $A_4$ and the holonomy of the 3d gauge field $A$ along the interval. Both are periodic scalars, so the target of the sigma-model is $T^2$. The BPS equations reduce to a holomorphic instanton equation $$ dA_3+\medstar dA_4=0, $$ which means that we are dealing with an A-model with target $T^2$. Its category of branes is the Fukaya-Floer category of $T^2$, which is fairly nontrivial (and by mirror symmetry equivalent to the bounded derived category of coherent sheaves on an elliptic curve). The A-model depends on the symplectic form on $T^2$ which can be read off the topological piece of the action (\ref{topaction}). Setting $A_1$ and $A_2$ to zero and reducing on an interval of length $2\pi$ it becomes $$ \frac{-4\pi^2 }{e^2}\int_{M_2} dA_3\wedge dA_4 $$ We may regard this expression as an integral of the pull-back of a symplectic 2-form $$ \frac{4\pi^2}{e^2} dx\wedge dy. $$ on the 2-torus with periodic coordinates $x,y$, both with period one. The symplectic area of this 2-torus is $4\pi^2/e^2$. We do not know how to describe the monoidal structure on this category arising from the fusion of boundary line operators. \subsection{The matter sector} As for $t=1$, we may consider either the Dirichlet or Neumann conditions for the scalars $\phi_3$ and $\phi_4$ (BRST-invariance requires them to be of the same type). In the Dirichlet case reduction on an interval gives the modified A-model whose only bosonic field is a real 1-form $\phi$ in two dimensions. As discussed above, it category of branes is the same as for a trivial TFT, i.e. it is equivalent to $D^b(Coh(\bullet))$. Unlike in the $t=1$ case, there are no ``winding sectors,'' since the scalars $\phi_3$ and $\phi_4$ are not periodic. So the category of boundary line operators in this case is $D^b(Coh(\bullet))$, with its standard monoidal structure. If $\phi_3$ and $\phi_4$ satisfy the Neumann condition, then the restriction of the 1-form $\phi$ to the 2d boundary must vanish. Reducing on an interval, we get an A-model whose only bosonic fields are $\phi_3$ and $\phi_4$, namely an A-model with target $T^*{\mathbb R}$. Its category of branes is the Fukaya-Floer category of $T^*{\mathbb R}$. Since this should be thought as the category of boundary line operators in a 3d TFT, it should have a monoidal structure. Since the only difference compared to the $t=1$ matter sector is the noncompactness of $\phi_4$, we expect that after we apply the equivalence of \cite{Nadler}, this monoidal structure becomes the standard monoidal structure on the constructible derived category of ${\mathbb R}$. \subsection{Putting the sectors together} \subsubsection{The DD condition} The DD boundary condition corresponds to a surface operator such that the 1-form $\phi$ has a fixed singularity of the form $$ \beta \frac{dr}{r}-\gamma d\theta, $$ while the holonomy of the gauge field $A$ is allowed to fluctuate, and the scalar field $\sigma$ vanishes at the insertion surface. To define such an operator properly, one has to excise a tubular neighborhood of the insertion surface and impose suitable conditions on the newly created boundary. Since the matter sector in the Dirichlet case does not have interesting boundary conditions, the category of boundary line operators is the same as in the gauge sector, i.e. the Fukaya-Floer category of $T^2$ with the symplectic area ${\mathfrak S}=4\pi^2/e^2$. From the 4d viewpoint, this is the category of line operators on the surface operator. Electric-magnetic duality maps the DD condition to itself. Indeed, it does not affect the matter sector, while in the gauge sector it maps the periodic scalar $A_4$ into a gauge field and maps the gauge field to a periodic scalar. Since in the DD case $A_4$ satisifies the Neumann condition, the dual gauge field satisfies the Dirichlet condition. Contrariwise, the Dirichlet condition for the gauge field is mapped by duality to the Dirichlet condition for the new periodic scalar. The only effect of duality is to replace $e^2$ with $4\pi^2/e^2$. Therefore the symplectic area of the $T^2$ is also inverted: $$ {\mathfrak S}\mapsto {\mathfrak S}'=\frac{4\pi^2}{{\mathfrak S}} $$ The Fukaya-Floer categories of two tori whose symplectic areas are related as above are equivalent by the usual T-duality. Moreover, we expect that the monoidal structure (which we have not determined!) is preserved by T-duality. \subsubsection{The NN condition} The NN condition corresponds to the surface operator such that $A$ has a fixed singularity of the form $$ \alpha d\theta, $$ while the singularity for the 1-form $\phi$ is allowed to fluctuate. To define such a surface operator properly, one has to impose suitable conditions on a boundary of a tubular neighborhood of the insertion surface. Upon reduction on an interval with NN boundary conditions on both ends, we get a 2d TFT which is a product of an A-type 2d gauge theory and an A-model with target $T^*{\mathbb R}$. Its category of branes is an equivariant version of the Fukaya-Floer category of $T^*{\mathbb R}$. It was conjectured above that it is equivalent to the equivariant constructible derived category of ${\mathbb R}$, with the standard monoidal structure (derived tensor product). Electric-magnetic duality maps the NN condition to itself, for the same reason as in the DD case. It acts trivially on the category of line operators, because the bosonic fields which survive the reduction on an interval (that is, $\sigma$, $\phi_3$ and $\phi_4$) are not involved in the duality. \subsubsection{The DN condition} The DN condition corresponds to a surface operator such that both $A$ and $\phi$ are allowed to have fluctuating singularities, while $\sigma$ has to vanish at the surface operator. Upon reduction on an interval with DN boundary conditions on both ends, we get a product of an A-model with target $T^2$ and an A-model with target $T^*{\mathbb R}$. Its category of branes is the Fukaya-Floer category of $T^2\times T^*{\mathbb R}$. Electric-magnetic duality maps the DN condition to itself. Its action on the category of line operators amounts to a T-duality on $T^2$ (duality acts trivially on the matter sector). The monoidal structure (which we have not determined) must be preserved by T-duality. \subsubsection{The ND condition} This case corresponds to the Gukov-Witten surface operator where the holonomy of $A$ is fixed, and the 1-form $\phi$ has a fixed singularity of the form $$ \beta \frac{dr}{r}-\gamma d\theta. $$ Reduction on an interval with ND boundary conditions gives a 2d TFT which is a product of an A-type 2d gauge theory and a modified A-model whose only bosonic field is a real 1-form. Since there are no interesting boundary conditions in the latter theory, the category of boundary conditions in this case is the same as in the former theory. That is, it is the $U(1)$-equivariant constructible derived category of sheaves over a point, or equivalently $D^b(Coh({\mathbb C}[2]))$ \cite{BL}. This is therefore the category of line operators sitting on the Gukov-Witten surface operator. The monoidal structure is the standard one (derived tensor product). In particular, since the trivial surface operator is a special case of the Gukov-Witten surface operator, we conclude that the category of bulk line operators in the GL-twisted theory at $t=0$ is $D^b(Coh({\mathbb C}[2]))$. In 4d terms, this can be interpreted as saying that all bulk line operators can be constructed by taking a sum of several copies of the trivial line operator and deforming it using the descendants of the BRST-invariant field $\sigma$ and its powers. This agrees with the results of \cite{Kap:qGL}, where it was argued that neither Wilson nor 't Hooft line operators are allowed at $t=0$. Electric-magnetic duality maps the ND condition to itself. It acts trivially on the category of line operators since the field $\sigma$ is not involved in the duality. \section{Conclusions} We have seen that GL-twisted gauge theory has a large number of surface operators other than the Gukov-Witten surface operators. These surface operators can be organized into a 2-category, and in the case $t=i$ we also proposed a description of this 2-category in terms of module categories. For $G=U(1)$ we proposed a similar description at $t=1$. It would be very interesting to find a physically-motivated\footnote{It was proposed by D. Gaitsgory that for $t\neq \pm i$ this 2-category can be described in terms categories with a D-module action of the loop group of $G$, but it is not clear how this proposal is related to the physical picture.} description of the 2-category of surface operators for all $t$ and $G$. Montonen-Olive duality implies that the 2-category of surface operators at $t=i$ in a theory with gauge group $G$ is equivalent to the 2-category of surface operators at $t=1$ in a theory with gauge group ${{}^L G}$. Moreover, these 2-categories both have braided monoidal structure, and the equivalence must be compatible with them. The usual statement about the equivalence of the categories of Wilson and 't Hooft line operators in the two theories follows from this. Indeed, bulk line operators can be regarded as endomorphisms of the trivial surface operator, and so must be equivalent (as tensor categories). From the mathematical viewpoint, the statement about the equivalence of braided monoidal 2-categories can be regarded as a 2-categorification of the geometric Satake correspondence.\footnote{The original version of the geometric Satake correspondence is due to Lusztig \cite{Lusztig} and can be regarded as a statement about the K-theory of the category of bulk line operators in the 4d TFT. That is, it is a statement about the commutative algebra which the 4d TFT attaches to $S^2\times S^1$. A way to categorify it to replace $S^2\times S^1$ with $S^2$; this corresponds to studying the symmetric monoidal category of line operators in the 4d TFT. This version of the geometric Satake correspondence has been proved by Ginzburg \cite{Ginz} and Mirkovic and Vilonen \cite{MV}. Alternatively, we obtain a 2-categorification of the geometric Satake correspondence by replacing $S^2\times S^1$ with $S^1$. Physically this corresponds to studying the braided monoidal category of surface operators in the 4d TFT.} The 2-category of surface operators at $t=0$ is a natural setting for studying local quantum geometric Langlands. In this case Montonen-Olive duality should give a nontrivial equivalence of braided monoidal 2-categories of surface operators in theories whose gauge couplings are inversely related. Already in the abelian case we saw that this equivalence is fairly nontrivial and reduces to T-duality in some special cases. As for the global quantum geometric Langlands, it seems natural to study the 2d TFT obtained by compactifying the 4d TFT at $t=0$ on a Riemann surface $C$ with an insertion of a surface operator of type DN or DD. This means that we cut a hole in $C$ and impose a boundary condition which allows the holonomy of $A$ along the boundary of the hole to be arbitrary. The resulting effective 2d TFT will have a category of branes which is a module category over the monoidal category of surface line operators of type DN or DD. The Montonen-Olive duality implies that replacing the group $G$ by its Langlands dual and inverting the gauge coupling gives rise to an equivalence of monoidal categories of surface line operators and a compatible equivalence of the categories of branes. We described the category of bulk line operators in the 4d theory at $t=i$, for a general gauge group. We found that it is equivalent to the equivariant derived category of coherent sheaves on the Lie algebra of the gauge group, with the linear coordinates on the Lie algebra sitting in cohomological degree $2$. This is a much larger category than one might naively expect based on special examples such as Wilson line operators. Finally, we showed that for $t=0$ and abelian gauge group the category of bulk line operators is fairly small (equivalent to $D^b(Coh({\mathbb C}[2]))$), and that electric-magnetic duality acts trivially on it. This agrees with \cite{Kap:qGL}, where it was shown that the $t=0$ theory does not admit either Wilson or 't Hooft line operators. It seems plausible that for a general gauge group electric-magnetic duality acts trivially on the category of bulk line operators at $t=0$.
\section{Introduction} There is a number of evidences indicating a deep relation between gravity and thermodynamics. In early 1970s four laws of black hole dynamics have been formulated \cite{Bardeen:1973gs}, whose form closely resembled the four laws of thermodynamics. It was then realized that this similarity between gravity and thermodynamics reaches far beyond formal analogy: the bold conjecture of Bekenstein \cite{Bekenstein:1973ur} that area of black hole horizon is proportional to thermodynamical entropy has been strengthened by Hawking discovery of black hole radiation \cite{Hawking:1974sw}. It turned out that indeed, as suggested by four laws of black hole dynamics, black holes behave as a thermal systems, with entropy and temperature proportional to the area and surface gravity, respectively. About twenty years later, in a remarkable paper Jacobson \cite{Jacobson:1995ab} has shown that from the proportionality between area and entropy \cite{Bekenstein:1973ur} taken as a fundamental principle one can derive the full Einstein equations of gravity. This idea has been then discussed in depth by Padmanabhan and others; see \cite{Padmanabhan:2009vy} for recent review and references. Building on these developments, in a recent paper \cite{Verlinde:2010hp}, Erik Verlinde argued that the force of the second law of dynamics and that of Newton's law of gravity can both have their origin in thermodynamics, and can be understood in terms of the entropic force (similar idea, based on equipartition of energy, has appeared earlier in \cite{Padmanabhan:2009kr}.) Within weeks several follow-up works appeared, testing this idea in various contexts (see, for example \cite{Li:2010cj} for the discussion in the context of cosmology and \cite{Wang:2010px} for derivation of the Coulomb law from thermodynamics.) In particular, in \cite{Smolin:2010kk} Smolin argued that Verlinde's proposal can be naturally realized in the context of Loop Quantum Gravity, and suggested its relations with constrained topological field theories. This idea is the starting point of the present work. Let us recall the major points of Verlinde's reasoning. The basic postulate of his work (see \cite{Verlinde:2010hp} for detailed discussion) is the following assumption \begin{itemize} \item Consider a holographic screen $\cal S$. If particle of mass $m$ crosses the screen, than the change of entropy of the screen is proportional to mass and displacement~$\Delta x$ \begin{equation}\label{1} \Delta S \sim m \Delta x\, . \end{equation} \item It then follows from the first law of thermodynamics that if there is the temperature $T$ that can be associated with the screen, then there exists the entropic force $F$ satisfying \begin{equation}\label{2} F \Delta x=T\Delta S\, , \end{equation} so that \begin{equation}\label{3} F \sim mT\, . \end{equation} \end{itemize} As shown by Verlinde the Newton's law of gravity can be derived assuming just from this postulate, energy equipartition, and the holographic principle. The reasoning goes as follows. Consider a spherical screen $\cal S$ at the center of which a localized, static chunk of matter of mass $M$ is placed. Assume that the radius of the screen is much larger than the size of the chunk, so that we can assume spherical symmetry of the problem. The holographic principle says that the number of bits $N$ on the screen $\cal S$ is proportional to its area\footnote{In what follows I will use the units in which the velocity of light $c$, the Planck constant $\hbar$, and the Boltzmann constant $k_B$ are all equal 1.} \begin{equation}\label{4} N = \frac{A}{G} \, , \end{equation} which is essentially the statement that the screen $\cal S$ is made of pixels of Planck size (this is the point where, as pointed out in \cite{Smolin:2010kk}, Loop Quantum Gravity with its quantization of area operator \cite{Rovelli:1994ge} naturally fits.) To complete the derivation of Newton's law one has to assume the equipartition of energy on the screen, from which it follows the relation between energy $E=M$ and the temperature \begin{equation}\label{5} M=\frac12\, NT\, . \end{equation} Finally, assume that the area of the screen $\cal S$ is \begin{equation}\label{6} A = 4\pi R^2\, . \end{equation} From equations (\ref{4}--\ref{6}) it follows that the temperature satisfies \begin{equation}\label{7} T = \frac{2GM}{4\pi R^2}\, , \end{equation} which, together with the postulate (\ref{3}) reproduces the Newton's law, $F=GMm/R^2$. It is worth noticing that for Schwarzschild black hole horizon $R=2GM$, eqn.\ (\ref{7}) reproduces the correct expression for Bekenstein--Hawking temperature $T_{BH}=(8\pi GM)^{-1}$. As argued by Verlinde the above reasoning is robust and general, the only weak point being the origin of the entropic force (\ref{2}), (\ref{3}). Certainly, there must be some microscopic degrees of freedom responsible for its emergence, and below I will argue that they, and the corresponding force, arise quite naturally in the formulation of gravity as a constrained topological BF theory. The plan of this note is as follows. In the next section I will recall the formulation of gravity as a constrained $\sf{SO}(4,1)$ BF theory and its coupling to particles. These technical results will be needed for the derivation of Verlinde's entropic force. The reader might decide to skip these technicalities and jump directly to Section III, where the main argument of the paper will be presented. The last section is devoted to discussion and conclusions. \section{Gravity as a constrained topological field theory} It is well known for quite some time that gravity can be formulated as a constrained topological field theory. The most popular popular model of this kind is given by Plebanski action \cite{Plebanski:1977zz}, being an action of the constrained BF theory of Lorentz $\sf{SO}(3,1)$ group. This model is a starting point for four dimensional spin foam models building (see e.g., \cite{Perez:2003vx}.) In the present context, however, it will be convenient to consider a different model, based on de Sitter gauge group $\sf{SO}(4,1)$ (the anti de Sitter model can be constructed analogously.) The main reason for this choice is that the $\sf{SO}(4,1)$ model allows for natural particles coupling. The action of the $\sf{SO}(4,1)$ constrained BF theory has the following form \cite{Smolin:2003qu}, \cite{Freidel:2005ak} \begin{equation}\label{8} S=\int \mathsf{B}^{IJ} \wedge \mathsf{F}_{IJ} -\frac{\beta}{2} \mathsf{B}^{IJ} \wedge \mathsf{B}_{IJ} -\frac{\alpha}{4} \mathsf{B}^{ab} \wedge \mathsf{B}^{cd} \epsilon_{abcd}\, , \end{equation} where $\mathsf{F}_{IJ}$ is the curvature of the $\sf{SO}(4,1)$ connection one-form $\mathsf{A}_{IJ}$ and $\mathsf{B}^{IJ}$ is a two-form valued in the algebra of the $\sf{SO}(4,1)$ group. Here the algebra indices $I, J, \ldots$ take values $0, \ldots, 4$, wile the indices $a,b,\ldots=0,\ldots, 3$ label Lorentz subalgebra $\sf{SO}(3,1)$ of $\sf{SO}(4,1)$. If one decomposes the connection $\mathsf{A}_{IJ}$ into Lorentz and translational parts \begin{equation}\label{9} \mathsf{A}^{ab}=\omega^{ab}, \quad \mathsf{A}^{a4} = \frac1\ell\, e^a \, , \end{equation} solves the equations of motion for $\mathsf{B}^{IJ}$ resulting from (\ref{8}) and plugs the result back to this action, as a result one gets the first order action of General Relativity $$ S= \frac{1}{2G}\int R^{ij}(\omega)\wedge e^k \wedge e^l \epsilon_{ijkl} - \frac{\Lambda}{12G}\int e^i \wedge e^j \wedge e^k \wedge e^l \epsilon_{ijkl} $$ accompanied by Holst term and a number of topological terms (see \cite{Freidel:2005ak} for details.) To get the action of General relativity the coupling constants $\alpha$, $\beta$ of the action (\ref{8}) and the length scale $\ell$ necessary for making the tetrad $e_\mu^a$ dimensionless are to be related to Newton's constant $G$, cosmological constant $\Lambda$, and Immirzi parameter $\gamma$ as follows \begin{equation}\label{10} \gamma=\frac{\beta}{\alpha}, \qquad \frac{1}{\ell^2}=\frac{\Lambda}{3}, \qquad G=\frac{3\alpha(1-\gamma^2)}{\Lambda} \end{equation} Let us pause for a moment to comment on the structure of the action (\ref{8}). If $\alpha$ vanishes the resulting action is just that of a topological field theory with no dynamical degrees of freedom. The local degrees of freedom of gravity (like gravitational waves or the presence of Newton's potential) appear only if the gauge breaking term, controlled by the coupling constant $\alpha$ is nonzero. Therefore the action (\ref{8}) clearly exhibits the split between topological and local degrees of freedom. In other words it is only the last term of (\ref{8}) that knows about dynamics of gravity. In the context of the present paper an obvious question arises: is it possible that the topological action describes the primary degrees of freedom of theory, while the gauge breaking term (i.e., gravity) arises as an emergent phenomenon from entropic force? As it is argued below at least Newton's force between massive bodies can be understood in this way. It is worth noticing also that the form of the gauge breaking term in the action (\ref{8}) is justified only by the fact that the theory described by this action turns out, at the end of the day, to be equivalent on shell to General Relativity. Therefore, it would be very interesting to find a principle, which would explain the presence of this term (see \cite{Smolin:2003qu} for an interesting proposal in this context.) Deducing Newton's law is the first step in this direction. As explained in \cite{Freidel:2006hv}, one can straightforwardly add point particles to the theory described by (\ref{8}) by identifying them with Wilson lines. To do that one includes the localized breaking of the gauge symmetry along the one dimensional world-line. The gauge degrees of freedom are then promoted to dynamical degree of freedom, which, in the case $\alpha\neq0$ reproduce the dynamics of a relativistic particle coupled to gravity. For a single particle this idea is realized by choosing a world-line $\cal P$ and a fixed element $\mathsf{K}$ in the Cartan subalgebra of the $\mathsf{so}(4,1)$ Lie algebra generated by two generators, the translational $T^{04}$ and rotational $T^{23}$ ones, depending on the particle rest mass and spin\footnote{Here we consider massive particles only. An extension to the case of massless particles is straightforward.} \begin{equation}\label{11} \mathsf{K}\equiv m \ell\, T^{04} + s T^{23} \end{equation} Note that the particle mass arises quite naturally in this picture in a purely algebraic way and is related to the one of the Casimirs of the gauge algebra. Then the action for the particle at rest takes the form \begin{equation}\label{12} S_{P}(\mathsf{A}) =-\int \mathrm{d} \tau\, \mbox{Tr} \left(\mathsf{K} \mathsf{A}_\tau(\tau)\right) \end{equation} where $\tau$ parametrizes the world line $z^\mu(\tau)$ and $\mathsf{A}_\tau(\tau) \equiv \mathsf{A}_\mu(z(\tau))\, \dot{z}^\mu$. The action of the particle moving in an arbitrary way is obtained by realizing that the moving particle is related to the one at rest by an appropriate $\sf{SO}(4,1)$ transformation acting on the world line. In this way the gauge degrees of freedom at the location of the particle become its physical degrees of freedom. Thus the Lagrangian of the dynamical particle has the form \begin{equation}\label{13} L(z,\mathsf{h}; \mathsf{A}) = -\mbox{Tr} \left(\mathsf{K} \mathsf{A}^\mathsf{h}_\tau(\tau)\right)\quad S = \int\, d\tau\, L(z,\mathsf{h}; \mathsf{A})\, , \end{equation} with $$ \mathsf{A}^\mathsf{h}=\mathsf{h}^{-1} \mathsf{A} \mathsf{h} + \mathsf{h}^{-1} d \mathsf{h}\, , $$ which can be rewritten as \begin{equation} L(z,\mathsf{h}; \mathsf{A})= L_1(z,\mathsf{h})-\mbox{Tr}(\mathsf{J} \mathsf{A}_\tau) \, , \label{14} \end{equation} with the first being the particle kinetic Lagrangian \begin{equation} L_1(z,\mathsf{h}) = -\mbox{Tr} (\mathsf{h}^{-1}\dot \mathsf{h} \mathsf{K})\, , \label{15} \end{equation} while the second describes its coupling to the connection $\mathsf{A}$, with $\mathsf{J}$ being the dynamical particle momentum/spin and is given by \begin{equation}\label{16} \mathsf{J} \equiv \mathsf{h}\, {\mathsf{K}} \,\mathsf{h}^{-1}\, . \end{equation} It can be shown that from (\ref{14}) the correct particle equation of motion (Mathisson--Papapetrou equation) follows; the theory described by (\ref{14}) and (\ref{8}) leads to Einstein-Cartan equations with point sources carrying mass and spin (see \cite{Freidel:2006hv} for detailed discussion.) This completes our description of the theory. Let us now turn to the discussion of solutions of topological BF theory coupled to such defined particle. Take the topological limit $\alpha\rightarrow0$ in (\ref{8}) and (\ref{14}) and consider the resulting field equations \cite{KowalskiGlikman:2006mu} for the particle at rest at the origin of an appropriate coordinate system\footnote{The reader may wonder that using the coordinates, and the geometry, we let gravity sneak through the back door. Of course, we need geometry to formulate the model, but the relation between local degrees of freedom of gravity and geometrical quantities is not present at this level yet, because dynamical gravity is not there.} One finds \begin{equation}\label{17} \mathsf{F}^{IJ} = \beta \mathsf{B}^{IJ} \end{equation} \begin{equation}\label{18} \mathsf{D}_\mathsf{A}\, \mathsf{B}^{IJ} = \mathsf{J}^{IJ}\, \delta_P, \quad \delta_P =\delta^3(x) \varepsilon \end{equation} where $\mathsf{D}_\mathsf{A}$ is covariant derivative of connection $\mathsf{A}$ and $\varepsilon$ is the volume three-form on a constant time surface. If one solves (\ref{17}) for $\mathsf{B}$ and substitutes the result to (\ref{18}) one finds that the left hand side of the resulting equation is zero by virtue of Bianchi identity. It is clear therefore that there does not exist a nonsingular connection $\mathsf{A}$ satisfying these equations for nonzero source. However, if one allows connections with string-like singularity (Misner string \cite{Misner:1963}, which is the gravitational counterpart of Dirac string) these equations can be solved. In fact it turns out that a pointlike source must be accompanied by a string extending from the source to infinity. As it was argued in \cite{KowalskiGlikman:2006mu} the spacetime corresponding to the solution of these equations\footnote{In the limit $\ell\rightarrow\infty$, which corresponds to the vanishing cosmological constant. Since the Taub-NUT charge $n$ does not depend on $\ell$, taking the limit does not influence it.} is the (linearized) Taub-NUT spacetime \begin{equation}\label{19} g = -\left(dt + n(1-\cos\theta)d\phi\right)^2 + dr^2 + r^2(d\theta^2 + \sin^2\theta\, d\phi^2) \end{equation} with Taub-NUT charge \begin{equation}\label{20} n=\bar G m, \quad \bar G = G\, \frac\gamma{1+\gamma^2}\, . \end{equation} This completes our brief description of constrained $\sf{SO}(4,1)$ BF theory, its relation to gravity and coupling to point sources. \section{Entropy and gravity from topological field theory} In the previous section I argued that if one couples the $\sf{SO}(4,1)$ topological BF theory (which after gauge breaking down to $\sf{SO}(3,1)$ is equivalent to General Relativity) to point particles, then the theory forces the particles to be accompanied by semi-infinite Misner strings. Moreover, the space-time corresponding to such solution is the Taub-NUT solution linearized in the charge $n$, which turns out to be proportional $Gm$, where the particle mass $m$ is the value of one of the Casimirs of $\sf{SO}(4,1)$ (the second one describes the spin, but here we discuss the spinless case only.) Knowing this let us turn to deducing the form of entropic force acting on the particle. Suppose the test particle of mass $m$ is at distance $R$ from the mass $M$, which we can assume to be also point-like. Consider now, as in Verlinde's argument, a spherical screen $\cal S$ of radius $R$. Let the test particle move radially towards the central mass piercing the screen, and let its displacement be $\Delta x$. As a result we have now a segment of the Misner string of the test particle of the length $\Delta x$ connecting it with the screen. Therefore the screen that previously was just a sphere\footnote{More precisely $\cal S$ consists of the sphere along with the attached string (or strings) emanating from the central mass $M$. But since we are only interested in the (infinitesimal) change of entropy, we do not have to consider them.} now becomes a sphere with a piece of Misner string, the line segment of length $\Delta x$ attached, $\cal S'$. Let me now turn to the main argument of this paper. It is well known that there is entropy associated with Misner string, see \cite{Hawking:1998jf}, \cite{Hawking:1998ct}, \cite{Carlip:1999cy}, and \cite{Mann:1999pc} where it is argued that the entropy of Misner string is intrinsically defined. In particular, using methods of conformal field theory Carlip \cite{Carlip:1999cy} shows that the segment of the Misner string of the length $\Delta x$ carries the entropy \begin{equation}\label{21} \Delta S = \frac1{8\pi G}\, n\, \Delta x= \frac1{8\pi }\, m\, \Delta x\, . \end{equation} Although this result has not been rigorously established in the present context of BF theory, it is unlikely that a formula analogous to (\ref{21}) does not hold in this case as well. It seems clear that Misner string carries entropy, no matter what is the theory describing local and/or topological degrees of freedom. If one accepts this argument, it follows from simple dimensional analysis that the entropy of the segment of Misner string of the length $\Delta x$ has to have the form \begin{equation}\label{22} \Delta S = \zeta\, m\, \Delta x\, , \end{equation} where $\zeta$ is the coefficient depending on the structure (and coupling constants) of the underlying theory. The entropy (\ref{22}) adds to the original entropy of the screen, and since it is proportional to the test particle displacement it leads to the emergence of the entropic force. Notice that since entropy increases when the test particle moves towards the mass $M$ this entropic force is attractive. Also when the test particle which was initially inside the screen moves outside, the entropy decreases, since the contribution from the Misner string is no longer present. Having (\ref{22}) it is possible now to run the remaining part of the Verlinde's argument essentially without modifications. The only point that is worth discussing is the equation relating the number of the screen pixels with area. Why $G$ is the measure of area of a pixel? In Loop Quantum Gravity this question finds its natural answer thanks to the fact that quantization of area in Planck scale units is the main result of this theory. It is not excluded that even in the context of BF theory one can define area operator with discrete spectrum. Until this idea is supported (or disproved) by concrete calculations we can rely only on general intuitions. The theory at hands provides us with the dimensionful scale $\ell$ and the dimensionless coupling constant $\beta$. From the two it is possible to construct another constant of dimension of area \begin{equation}\label{23} \bar G = \frac{3\beta}{\ell^2} \end{equation} which in the full theory (including nontrivial gauge breaking term) becomes proportional to Newton's constant of general relativity (cf.\ (\ref{10})). Since $\beta$ has some final value, and since $\ell$ is an infrared scale of the theory, it is quite natural to treat $\bar G$ as an intrinsic ultraviolet scale of the theory, and thus to replace (\ref{4}) with \begin{equation}\label{24} N = \frac{A}{\bar G}, \quad A=4\pi R^2\, , \end{equation} which directly, by virtue of Verlinde's argument recalled in Introduction leads to the Newton's law \begin{equation}\label{25} F=\frac{G\, m\, M}{R^2} \, , \end{equation} where $G=4\pi \zeta\bar G$ is the Newton's constant, whose value can be directly measured, e.g., in Cavendish experiment. This concludes the presentation of the main argument of this paper. \section{Conclusions and outlook} In this note I argued that the form of entropic force being the starting form of the recent proposal of Verlinde \cite{Verlinde:2010hp} to seek the origin of gravity in thermodynamics can be understood if one assumes that the fundamental degrees of freedom behind it are described by the topological BF theory coupled to particle(s). The reason for this is that, as shown in \cite{Freidel:2006hv} and discussed in \cite{KowalskiGlikman:2006mu}, a particle carrying the charge of (anti) de Sitter $\sf{SO}(4,1)$ ($\sf{SO}(3,2)$) group coupled to the topological BF theory with the same gauge group must have Misner string attached. This string, in turn, carries entropy, which adds to the entropy of the holographic screen $\cal S$ when the particle crosses it, which results in emergence of the entropic force. There are at least two problems that has to be solved before this idea becomes a solid proposal. First, one has to calculate the entropy of Misner string directly in the framework of BF theory, to fix the constant $\zeta$ in (\ref{22}). This can be presumably done with the help of the method similar to that described in \cite{Carlip:1999cy}. Second, it would be interesting to see if it is possible to improve on the part of the original Verlinde's argument that makes use of equipartition of energy to get the expression for temperature (\ref{7}). It would be desirable in this context to check if one can define for BF theory an area operator with discrete spectrum, resembling this of Loop Quantum Gravity \cite{Rovelli:1994ge}. The work on both these problems is in progress. \begin{acknowledgments} I would like to thank Lee Smolin for discussion and encouragement. This work is supported in part by research projects N202 081 32/1844 and NN202318534 and Polish Ministry of Science and Higher Education grant 182/N-QGG/2008/0. \end{acknowledgments}
\section{Introduction} \label{intro} \footnotetext[1]{This work was based on observations with the NASA/ESA Hubble Space Telescope, obtained at the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc.,under NASA contract NAS 5-26555. } The generic properties of globular clusters, which are generally considered to be free of significant amounts of dark matter, are different from those of the dark-matter-dominated dwarf galaxies. These classes of objects were found for a long time to occupy different regions of parameter space in terms of, e.g., scale size, luminosity, and internal velocity dispersion \citep[e.g.][and references therein]{mateo98}. New systematic surveys for globular clusters and dwarf satellites over the last few years have generated a shift in our understanding of the connection between those two classes of objects. On one hand, several of the newly discovered dwarf satellites around the Milky Way \citep[e.g.][]{willman05,belokurov07} and Andromeda \citep[e.g.][]{zucker04,martin06,irwin08} exhibit luminosities and/or sizes that are strikingly comparable to those of globular clusters. On the other hand, luminous star clusters with large sizes have been discovered around, e.g., M31 \citep{huxor05} and M33 \citep{stonkute08}. In contrast, in the Milky Way globular cluster system, most of the unusually extended star clusters are the Palomar-type objects that lie preferentially at large Galactocentric distances and are predominantly faint \citep[e.g.][]{vandenbergh04}. The extended luminous star clusters are found to resemble the classical globular clusters in terms of their stellar population properties, i.e., they are generally old and metal-poor, albeit with combined structures and luminosities unlike those observed for any other globular clusters in the Local Group or beyond \citep{mackey06}. They are found to occupy the gap in parameter space between classical (compact) globular clusters and dwarf spheroidal galaxies \citep{huxor05}. Globular clusters are typically found in much larger numbers within giant elliptical galaxies than in spirals such as the Local Group large galaxies. At a distance of 3.8 Mpc \citep{rejkuba04,harris09}, NGC~5128 is the closest giant elliptical. Its globular cluster system, the largest of any galaxy within ${\rm \sim 15\,Mpc}$, has been studied photometrically and spectroscopically for many years \citep[for only the most recent such work and references to earlier papers, see, e.g.][]{harris04,woodley07,gomez07,mcl08,woodley09}, and at present a total of 605 clusters have been individually identified. Extended (and bright) globular clusters are found to represent a tiny fraction of its overall star cluster population. \citet{gomez07} indicate that globular clusters with half-light radii exeeding 8\,pc make up a few percent of the their spectroscopically confirmed globular clusters. These extended globular clusters have luminosities comparable to those of extended clusters in the outskirts of M31 \citep[see Fig.~8 of][]{gomez06}), brighter than most Galactic globular clusters of similar sizes. The most massive clusters in NGC~5128 are found to follow a mass-size relation, and to have higher mass-to-light ratios on average than clusters with masses below ${\rm 10^{6} M_{\odot}}$ \citep{rejkuba07}. In this contribution, we report the discovery of two faint, extended globular cluster candidates in the outskirts of NGC~5128, one of which is a high-probability old globular cluster based on direct colour-magnitude photometry. The dataset used here was initially obtained to study the resolved stellar populations in the remote outskirts of NGC~5128, allowing us to investigate the properties of the neighborhood of the extended faint star cluster to shed light on its origin. The layout of this paper is as follows: Section \S~\ref{data} briefly describes the data set, while Section \S~\ref{results} presents the properties of the reported halo globular cluster. In \S~\ref{disc} we will discuss the present work and its implications. \S~\ref{summary} gives the summary. Throughout the paper we use an intrinsic distance modulus for NGC 5128 of$(m-M)_{\circ}=27.90$ following the recent synthesis of five standard-candle distance calibrations given by \citep{harris09}. The foreground reddening toward the field was estimated to be $E(B-V)=0.11$ from the all-sky map of \citet{schlegel98}. We have neglected the effect of any possible differential reddening across the targeted field and along the line of sight from within NGC 5128 itself, since the target field is one located in its outer halo. The differential reddening due to the Galactic foreground across the $3.4'$ width of an HST/ACS field at this Galactic latitude ($b = 19^o$) is also likely to be negligible. \section{Data} \label{data} The imaging data we use here is the single very deep pointing obtained by \citet{rejkuba05} with the Advanced Camera for Surveys (ACS/WFC) camera on the Hubble Space Telescope (HST) in program GO-9373. This field, imaged in $F606W$ (wide $V$) and $F814W$ ($I$), is located at a linear projected distance of approximately $\sim 38{\rm\,kpc}$ South of the NGC 5128 center. The target field was chosen to avoid any known surface brightness anomalies such as jet-induced star-forming regions \citep{mould00,rejkuba02}, shells and arcs \citep{malin83}, and dust lanes \citep{stickel04}. The total exposure consisted of of 12 full orbits in each filter, reaching $F606W \sim 30$, $F814W \sim 29$ and making this material by far the deepest photometric probe into the stellar population of NGC 5128. In terms of limiting absolute magnitude, it is deep enough to resolve horizontal-branch stars in the halo and to show the structure of the ``red clump'' (the core-helium-burning stars in the colour-magnitude diagram). The observations and reduction techniques are fully described in \citet{rejkuba05} and we therefore only briefly summarize this information here. The photometry of \citet{rejkuba05} was measured with the DAOPHOT suite of codes \citet{stetson1994}. From this we select stars with high quality measurements, including sharpness parameter $< 2$, goodness of fit $\chi^{2} < 3$, and $\sigma_{\rm F606W} < 0.3$. We further limited the catalogue to stars with $-0.5 < (V-I)_{\circ} < 4.0$. For an old simple stellar population, redder stars are more metal-rich than bluer ones, and the width of the red giant branch (RGB) at a given luminosity is a very direct indicator of the spread in the stellar metallicity. The analysis of the properties of the age-sensitive features of the asymptotic giant branch bump and the red clump of the stellar populations in the outskirts of NGC~5128 indicates that they are predominantly old. While we cannot rule out a modest age spread in these halo stars \citep[see][]{rejkuba05}, the colour spread of the RGB stars observed in the data is due primarily to a spread in metallicity at an old mean age and can be used to derive the halo metallicity distribution function (MDF) \citep[e.g.][and references therein]{harris99,harris00,harris02,mouhcine05}. To estimate the metallicities of the survey stars, we repeat the procedure explained in \citet{mouhcine07}, interpolating between the models of \citet{vandenberg06} of $\alpha$-enhanced red giant branch stars of mass $0.8\,M_{\odot}$. The models span ${\rm -2.314 < [Fe/H] < -0.397}$ approximately in steps of 0.1 dex, and are complemented at the metal rich end by two models with ${\rm [Fe/H] = 0.0}$ and ${\rm [Fe/H] = +0.4}$ (see \citealt{mouhcine05} and \citealt{harris00} for more details on the interpolation procedure and the calibration of the track grid). Stars outside of the colour-magnitude range covered by these RGB tracks were flagged and not used in subsequent analysis (i.e., no extrapolation beyond the validity of the models was attempted). The metallicity of stars on the red clump cannot be interpolated from RGB tracks; we therefore imposed a faint cut-off at $M_I < -0.75$ to retain only the brighter RGB stars, which also have the advantage of being clearly brighter than the completeness limits of the data. This limit corresponds to $I_{\circ} < 27.17$, or $I < 27.38$, where the $I$-band measurement uncertainty is $\sigma_I=0.13$. An additional quality cut of $\sigma_V=0.13$ was imposed to ensure that the $V$-band measurements were also of good quality. \section{Results} \label{results} \subsection{A new faint globular cluster} \begin{figure} \includegraphics[clip=,width=0.45\textwidth]{xy_map_feh.ps} \caption{Spatial distribution of stellar sources classified as RGB candidates over the observed field. Large solid points show the distribution of metal-poor, i.e., [Fe/H] $< -1.3$, stars. The small compact clump of metal-poor stars at top center shows the position of the candidate star cluster G0606. The cross indicates the position of the second star cluster candidate G0607. The position scales are normalized to $(0,0)$ at the cluster candidate G0606.} \label{xy_map} \end{figure} \begin{figure*} \includegraphics[angle=0]{clusters_pal14.eps} \caption{Thumbnail images in $V$ of the two cluster candidate, GC0606 (left panel) and GC0607 (center panel) (see the text for the adopted numbering scheme). Field size of each thumbnail is 4 arcsec across, equivalent to 75 pc at the distance of NGC 5128. Note the very bright star just off the right side of the frame for GC0606. The \emph{right panel} shows a thumbnail image for comparison of the Milky Way outer-halo cluster Palomar 14, drawn from the HST/ACS archives. Its luminosity and effective radius are similar to our two cluster candidates. To compensate for the $50\times$ nearer distance of Pal 14, its image was resampled with a Gaussian convolution and then block-averaged to the same linear size.} \label{images_gc} \end{figure*} The overwhelming majority of stars in this outer-halo pointing are the rather uniformly distributed halo field stars. However, the long dynamical timescales of regions this far out (38 kpc and above) allow the possibility that identifiable substructures, such as extended globular clusters (GCs) or faint satellite dwarfs might be present as well. An efficient way for revealing the presence of these stellar (sub-)structures in the presence of heavy field contamination, especially when integrating stellar populations along a line of sight, is to investigate differences in the spatial distribution of stars with different metallicities \citep{ibata07,ibata09}. Figure\,\ref{xy_map} shows the spatial distribution of RGB stars over the field. The sample of all the field stars is shown as small dots, while \emph{metal-poor} RGB stars, i.e., those with photometrically determined abundances [Fe/H] $< -1.3$ dex, are shown as larger solid dots. The most striking feature in the stellar spatial distribution is a small, compact clump of metal-poor RGB stars near the west side of the frame (see Fig.~\ref{xy_map}). No such clumps are found for the more chemically evolved stellar populations. To shed light on the nature of the identified concentration of metal-poor stars, we show on the left panel of Figure\,\ref{images_gc} a zoomed thumbnail image from the ACS $F606W$ stacked frame. The object clearly has the morphology of a faint, extended star cluster, more diffuse in structure than the classical globular clusters that have previously been observed in NGC~5128 (see, e.g. the sample images in \citet{harris06}). Following the homogeneous catalogue numbering system for NGC 5128 GCs defined by \citet{woodley07} and continued in \citet{woodley09}, we name this new GC candidate GC0606. This new object strongly resembles the Palomar-type clusters found in the Milky Way outer halo. For more direct visual comparison, the right panel of Figure\,\ref{images_gc} shows a thumbnail image of the Milky Way outer-halo cluster Palomar 14 (which is $\sim 70$\,kpc from the Sun), drawn from the HST/ACS archives. To compensate for the difference in distances between NGC~5128 and Pal 14, i.e., a factor of 50, the Pal 14 image was resampled with a Gaussian convolution and then block-averaged to the same linear size. Pal 14 has a scale size (half-light radius 24 pc) and luminosity ($M_V = -4.7$) that are similar to our NGC 5128 candidates (see below), demonstrating that these faint, extended GCs can be discovered readily with this technique. \begin{figure} \includegraphics[clip=,width=0.45\textwidth]{cmd_bright_object.ps} \caption{The colour-magnitude diagram of stars lying within a radius of 2 arcsec from the centre of the cluster candidate GC0606. Shown also as solid lines are the red giant sequences of the indicated Galactic globular clusters from \citet{dacosta90} assuming an intrinsic distance modulus for NGC~5128 of 27.9. A sequence of red giant stars is clearly visible. } \label{gc_cmd_fidu} \end{figure} \begin{figure} \includegraphics[clip=,width=0.45\textwidth]{gc_cenA_obs_vs_basti.ps} \caption{The cluster colour-magnitude diagram of the cluster candidate GC0606 compared to red giant branches from scaled-solar (the left panels) and $\alpha$-enhanced BaSTI isochrones with ${\rm [\alpha/Fe]=+0.4}$ (the right panels). The upper panels show the effect of varying the age at a fixed overall metallicity, while the lower panels show the effect of varying the overall metallicity at a fixed age. } \label{gc_cmd_basti} \end{figure} A powerful tool to investigate the nature of the candidate is the colour-magnitude diagram (CMD) of its stars. A classical globular cluster with a half-light radius of a few parsecs would subtend only a small angle: a typical half-light radius of 3\,pc converts to 3 ACS pixels, not much larger than the 2-px Full Width at Half Maximum of the stellar point spread function. The outer parts of normal, luminous GCs will have enough stars that a CMD can be obtained; this was first done for NGC 5128 by \citet{harris98} for the cluster C44=GC0227, an inner-halo cluster which turned out to have moderately low metallicity from its RGB locus. The deep WFPC2/PC1 $(V,I)$ images used for their study reached to $I(lim) \simeq 27$, not as deep as our ACS data. In the case of faint and diffuse star clusters, the situation is less desperate. Even though the total RGB population within the cluster GC0606 is low, the crowding is not extreme and the ACS images contain a number of resolved members for which the PSF fitting procedure allows accurate photometric measurements. In order to place constraints on the stellar population properties of GC0606, we show in Fig.\,\ref{gc_cmd_fidu} the CMD of all stellar objects detected within a circular region with a radius of two arcseconds (40 px) around the star cluster. Fiducials of RGB sequences for Galactic globular clusters spanning a wide range of metallicities from \citet{dacosta90} and shifted to the distance of NGC~5128 are overplotted. The brightest stars in the cluster are nicely aligned with the red-giant sequences at intermediate metallicity, between the fiducial sequences for ${\rm [Fe/H]\sim -1.3}$ and $-1.55$, particularly over the upper two magnitudes where the photometry is of the highest quality ($\sigma_{V,I} \la 0.05$). This CMD not only confirms the identity of GC0606 as an old globular cluster, but also places it comfortably within the metal-poor sub-population of the bimodal metallicity distribution function of the globular cluster system of NGC~5128 that peaks at ${\rm [Fe/H]\sim-1.5}$ \citep[e.g][]{harris04,beasley08}. To strengthen the constraints on the properties of the star cluster, we compare the stellar photometry to the BaSTI theoretical isochrones for both scaled-solar and $\alpha$-enhanced compositions \citep{pietrinferni04,pietrinferni06}. Fig.\,\ref{gc_cmd_basti} shows aligned Solar-scaled (left panels) and $\alpha$-enhanced (right panels) isochrones on the star cluster CMD. The upper panels show comparison with isochrones with ages ranging from 6\,Gyr to 13\,Gyr for an overall metallicity of ${\rm [M/H]=-0.96}$, while the lower panels show 10\,Gyr isochrones with metallicities ranging from -0.66 to -1.5. Considering first the Solar-scaled isochrones, those with a metallicity ${\rm [M/H]=-0.96}$ and ages older than 6\,Gyr are found to provide the best fit along the majority of the RGB. The BaSTI theoretical isochrones also give a metallicity estimate in good agreement with our empirical value of ${\rm [Fe/H]\sim -1.3}$, keeping in mind that metal-poor GCs typically exhibit enhancement in $\alpha$-elements of approximately ${\rm [\alpha/Fe] \sim +0.3}$ \citep[see e.g.][]{carney96}. The comparison of the star cluster CMD to the BaSTI isochrones with an $\alpha$-enhanced composition, i.e., ${\rm [\alpha/Fe] = +0.4}$, indicates that the isochrones with ${\rm [M/H]=-0.96}$ and ages older than ${\rm \sim\,6\,Gyr}$ provide again the best fit to the cluster RGB. Once again, this is in excellent agreement with the metallicity estimate obtained by comparing the star cluster to Galactic cluster fiducials. \subsection{Structural Parameters} To characterize further the globular cluster, we measured its overall structural parameters. A surface brightness profile of the globular cluster was generated by carrying out aperture photometry of the cluster on both the $V$ and $I$ master images. We derived the profile from \emph{daophot/apphot} direct concentric-aperture photometry, using the flux differences between apertures of adjacent radii to derive the surface brightness in successive annuli. A problem we faced in this analysis was the presence of a nearby very bright star, which prevented the profile from being measured further out than 16 px radius ($0.8''$) in $I$ and 26 px ($13''$) in $V$. The profile fitting code of \citet{mcl08} was then used to fit \citet{king62} and \citet{king66} profiles, numerically convolved with the stellar PSF as defined from nearby bright, uncrowded stars on the frames. The best-fit results for both model profiles are shown in Figure \ref{brightprofile}. As is evident from the figure, the cluster is very much more extended in characteristic size than the PSF, so the fitting results are insensitive to the details of the PSF or any small changes it might have across the field. The center coordinates are no more certain than $\pm 2$ px ($0.1''$) on each axis. The cluster is almost entirely dominated by a large, flat central core. In the sense defined by the King models, the cluster has an extremely low formal value of the central concentration $c = {\rm log} (r_t/r_c)$ and central dimensionless potential $W_0$, making it similar to the lowest-concentration globular clusters known. The half-light radius $r_h$ was determined by numerically integrating the PSF-deconvolved King profile solutions out to large radius and then finding the projected radius enclosing half the total luminosity. Lastly, we estimated central surface brightness, which is also insensitive to the details of either the profile fit or the PSF because the cluster has essentially constant surface brightness within 5 px. All the integrated magnitudes and central surface brightness values are internally uncertain to at least $\pm 0.1$ magnitude and perhaps more. The compiled measured parameters are shown in Table \ref{list_gc_candidates}. The apparent magnitudes and colours have been converted to luminosities and intrinsic colours assuming $(m-M)_I = 28.10$ and $E_{V-I} = 0.14$, and conversion of half-light radii to parsecs assumes $d = 3.8$ Mpc as stated earlier. It is evident that the cluster is just as diffuse and low-luminosity as its first visual impression gives. \begin{table} \caption{Physical Parameters of Faint Extended Globular Cluster Candidates} \label{list_gc_candidates} \begin{tabular}{lcc} \hline Parameter & GC0606 & GC0607 \\ \hline $\alpha$ (J2000) & 13:25:12.40 & 13:25:19.78 \\ $\delta$ (J2000) & -43:35:15.44 & -43:34:40.30 \\ $I$(tot) & 21.9 & 24.0 \\ $(V-I)$ & 1.15 & 0.84 \\ $\mu_0(I)$ (mag arcsec$^{-2}$) & 23.0 & 24.1 \\ $r_h$ (arcsec) & $0.80 \pm 0.03$ & $0.47 \pm 0.03$ \\ $M_I$ & -6.2 & -4.1 \\ $M_V$ & -5.2 & -3.4 \\ $r_h$ (pc) & 14.7 & 8.7 \\ $L_V/L_{\odot}$ & $10.3 \times 10^3$ & $1.9 \times 10^3$ \\ \end{tabular} \end{table} \begin{figure} \includegraphics[angle=0,width=0.5\textwidth]{brightprofile.eps} \caption{Surface brightness profile for the brighter candidate, GC0606. The measurements were made through direct concentric-aperture photometry as described in the text. Solid dots show the measurements in concentric annuli with internal error bars, while open circles show the fiducial profile for the stellar PSF. The best-fit King (1966) model is shown as the solid line, while the King (1962) model fit is shown as the dashed line. Left and right panels show the $I-$band and $V-$band results separately; the vertical axis give the surface brightness normalized to the central value $\mu_0$. } \label{brightprofile} \end{figure} \begin{figure*} \includegraphics[angle=0,width=0.45\textwidth]{cmd_faint_object.ps} \includegraphics[angle=0,width=0.45\textwidth]{faintprofile.eps} \caption{Left: The colour-magnitude diagram of stars lying within a radius of 2 arcsec from the centre of the faint cluster GC0607. Shown also as solid lines are the red giant sequences of the indicated Galactic globular clusters from \citet{dacosta90}. Right: Surface brightness profile for the fainter of the two cluster candidates, GC0607, measured with ellipse-fitting as described in the text. Symbols and lines are the same as in Fig.\,\ref{brightprofile}.} \label{faintcluster} \end{figure*} \subsection{A second extended cluster candidate} Motivated by the discovery of the candidate GC0606, we visually examined more closely the $V$ and $I$ master images looking for even fainter objects that could plausibly be Palomar-like GC candidates. A candidate located at $\alpha = 13^{\rm h}25^{\rm m}19.78^{\rm s}$, $\delta = -43\degr 34\arcmin 40.30\arcsec$ was found, which we label as GC0607. The middle panel of Figure\,\ref{images_gc} shows a zoom on the vicinity of this object. This second candidate draws attention to itself primarily as a symmetric, low-surface-brightness patch that is only marginally resolved into a few stars; it has few or no obvious bright RGB stars that would have allowed it to be picked up through our initial search technique (Fig.~\ref{xy_map}). The left panel of Fig.\,\ref{faintcluster} shows the CMD of all stellar objects detected within a circular region of radius of two arcminutes around the star cluster candidate. RGB fiducials are overplotted as before. No clear sequence of RGB stars is present, but the CMD contains a few stellar objects with luminosities and colours consistent with those expected for horizontal branch stars at the distance of NGC~5128. For comparison, the handful of faintest known GCs in the Milky Way such as Pal 13, Pal 1, or AM4 (all of which lie at $M_V < -4$ integrated luminosity) are almost entirely lacking in RGB stars brighter than the horizontal branch, to such an extent that the addition or subtraction of just one RGB star would change the total cluster luminosity very noticeably. To determine the structural parameters of the cluster candidate, which is (mostly) unresolved, we used \emph{stsdas/ellipse} to construct the empirical radial profile of light intensity out to a radius of 30 px ($1.5''$). Note that the circular aperture photometry leads to the same results to within the errors. As before, the comparison profile of the stellar PSF was measured from adjacent bright, unsaturated, uncrowded stars. The profile fitting code of \citet{mcl08} was then used to fit \citet{king62} and \citet{king66} profiles, numerically convolved with the PSF. The best-fit results are shown in the right panel of Figure \ref{faintcluster}, with the measured parameters listed in Table \ref{list_gc_candidates}. As for the first extended cluster, GC0607 is almost entirely dominated by a large, flat central core. Again, the central concentration and central dimensionless potential are similar to the lowest-concentration globular clusters known in the Milky Way. The half-light radius, the central surface brightness, and the total luminosity in each photometric band were determined in the same manner as for the first faint extended globular cluster. The difference in $(V-I)$ colour between GC0606 and 0607 is surprisingly large, but can be at least partly attributed to the relative dominance (or lack) of RGB stars brighter than the horizontal branch. GC0606 has numerous RGB stars, as seen from the colour-magnitude diagram, while GC0607 has virtually none and so its colour (if it is indeed a GC) would be dominated by the bluer subgiant and turnoff stars. The integrated luminosity, colour, and scale size of GC0607 are all consistent with it being an extremely faint, diffuse GC. However, further and more definitive confirmation (such as by radial velocity measurement) will be exceptionally hard to obtain. Although we suggest that GC0607 should be kept as a plausible cluster candidate, the possibility cannot be ruled out that it is a faint background galaxy. \begin{figure} \includegraphics[clip=,width=0.45\textwidth]{rh_mv_muv.ps} \caption{Half-light radius (in pc) versus V-band luminosity for globular clusters in nearby galaxies. The Milky Way GCs from \citet{harris96} are shown as solid dots. Those with the central surface brightness values larger than ${\rm 23 mag\,arcsec^{-2}}$ are encircled. GCs of M31 \citep{barmby07} and NGC~5128 \citet{gomez06,mcl08} are shown as asterisks and open circles respectively. M31 extended clusters from \citep{mackey06} are shown as solid squares. Open and solid pentagons show respectively the extended clusters of M33 \citep{stonkute08}, and Scl-dE \citep{dacosta09}. The extended clusters discussed in the paper are shown as open stars. The solid line shows the relation $\log(r_h)=0.2\times M_V + 2.6$ from \citet{vandenbergh04}, and the dashed line represents a constant surface luminosity of ${\rm 15\,L_{V,\odot} \times pc^{-2}}$ within the half-light radius. The Galaxy GC $\omega$ Cen and NGC 2419 are labeled.} \label{rhmv} \end{figure} \begin{figure} \includegraphics[clip=,width=0.45\textwidth]{median_feh_map_CenA_40kpc.ps} \caption{Two-dimensional median metallicity map over the observed ACS field. Stars shown in Fig.~\ref{xy_map} binned into a $12\times12$ super-pixel array. } \label{map_median_feh} \end{figure} \section{Discussion} \label{disc} Fig.\,\ref{rhmv} shows the location of the two new GC candidates we have discovered here on the structural parameter plane of half-light radius versus V-band luminosity, compared to those of the Milky Way \citep{harris96}, M31 \citep{barmby07} and NGC~5128 \citep{gomez06,mcl08} shown as solid dots, asterisks, and open circles respectively. Galactic GCs with V-band extinction-corrected central surface brightness values fainter than ${\rm 23\,mag\,arcsec^{-2}}$ are encircled. The extended clusters reported around M31 \citep{mackey06}, M33 \citep{stonkute08}, and the Sculptor Group dwarf elliptical Scl-dE1 \citep{dacosta09} are also shown. The solid line shows the equation derived by \citet{vandenbergh04} who found that only two GCs lie above this line. A constant surface luminosity of ${\rm 15\,L_{V,\odot} \times pc^{-2}}$ within the half-light radius is also shown as the dashed line. This line seperates nicely the subpopulation of extended clusters from the rest of the GC population. The GC candidates we found in NGC~5128 appear different from the populations of extended clusters recently reported in the literature. Those objects are bright, i.e., $M_{V,\circ} \la -6.5$, while the extended clusters of NGC~5128 are much fainter matching up well, particularly the G0606 cluster, with the faint and extended cluster population in the Milky Way. Interestingly, the group of Galactic GCs with luminosities and half-light radii comparable to those of G0606 are situated all at galacto-centric distances $\ga 25\,$kpc. Only one Galactic GC beyond 25\,kpc, i.e., Palomar 13, is not in this group of clusters. It is worth to mention that the central surface brightness of the object of the group of Galactic GCs with luminosities and half-light radii similar to those of G0606 that is not encircled, i.e., Pyxis, is not available. By examining the distribution of old globular clusters in a number of different galactic systems in the size vs. luminosity diagram, \citet{dacosta09} have argued that the distribution function of globular cluster sizes is most likely bimodal. A first dominant subpopulation has half-light radii peaking at $\sim3.0\,$pc, but there is a secondary subpopulation with half-light radii peaking at $\sim10\,$pc and a dearth of GCs with half-light radii around 5\,pc. They suggested that this indicates the presence of two distinct modes of star cluster formation, with the less common extended clusters being primarily formed in the gravitationally smoother environment of dwarf galaxies compared to larger galaxies. An intriguing implication of this correlation is that extended and diffuse GCs in large galaxies, which tend to be found predominantly in their outer regions \citet{vandenbergh04,huxor05,huxor09}, may be of accretion origin. By investigating the distributions of the Milky Way globular cluster properties, their spatial distribution, and by comparing with those of globular clusters in dwarf galaxies, \citet{vandenbergh04} have suggested indeed that the faint extended globular clusters in the outer Galactic halo originate most likely from the disruption of now defunct dwarf companions. Could the diffuse and under-luminous globular cluster we have discovered in the outskirts of NGC~5128 be the counterpart of the potentially accreted Galactic outer halo extended clusters? If a star cluster is of accretion origin, one would expect, if the accretion event is not too old and/or the accreted stars are not fully mixed yet, that the diffuse stellar populations in its close vicinity would be different from the average halo stellar populations. The depth of our ACS imaging data resolving the stellar content of the halo down to the core helium burning stars gives us the opportunity to investigate the properties of the diffuse stellar populations in the immediate neighborhood of the extended globular cluster. The ACS image contains many large background galaxies, as well as several bright stars (see Fig.\,1 of \citet{rejkuba05}) whose ``halos'' cause localised holes in the star-counts map. A mask was constructed by choosing suitably large elliptical areas around these problematic regions. We have binned the stars in a $12\times12$ grid in order for each super-pixel to contain $\sim 100$ sources (to have signal to noise ratio S/N$\sim 10$). Each superpixel is thus $17''$ wide or 314 parsecs projected width. Only stars outside of the masked regions were kept. A detailed and quantitative analysis of the spatial sub-structures over the ACS field is deferred to a subsequent contribution, but it is worth mentioning here that the stellar surface density of RGB stars over the field shows no obvious coherent stellar structure that might indicate the presence of streams or shells of stellar debris. Figure \ref{map_median_feh} shows the two-dimensional map of the median values of metallicity of RGB stars calculated in the super-pixel grid. For reference, the extended globular cluster discussed in the present paper is located within the most metal-poor super-pixel, the dark blue spot of Figure \ref{map_median_feh}. The random errors on the median metallicities in the super-pixels, due to both photometric uncertainties and population sampling effects, estimated as described in full details in \citet{ibata09}, are small, i.e., typically $\sim 0.02$~dex. A striking feature of the median stellar metallicity map is the large pixel-to-pixel variation, much larger than the typical random uncertainties. It is worth noting that those small-scale chemical variations have no obvious correspondence in the stellar density map. Due to the virtually negligible contamination from foreground or background sources, the present observations are extremely sensitive to the presence of sub-structures. It is in principle possible with the ACS to detect a population of $\sim 15$ RGB stars scattered over a volume of many hundred cubic kpc, making it much easier to detect sub-structures from spatial metallicity variations rather than from the enhancement they cause in the stellar density map \citep[see][for more details]{ibata09}. The outer regions of NGC~5128 have been found to be predominantly populated by moderately metal-rich stars, with an average metallicity of ${\rm [m/H]\sim -0.5}$ and only $\sim 10\%$ of the stellar populations more metal-poor than ${\rm [m/H]\approx -1}$ \citep{harris00,harris02,rejkuba05}. A gradient of the median metallicity is apparent in the median metallicity map, with the top half of the ACS field being more metal-poor than the lower half. As discussed at length by \citet{rejkuba05}, the metallicity distribution function of the overall stellar populations at $\sim 40$\,kpc is very similar in term of its shape, average metallicity, and metallicity dispersion, to those determined at $\sim 20$ and $\sim 30$ kpc respectively from the centre of the galaxy \citep{harris99,harris00}. No detectable radial metallicity gradient is present in the outskirts of NGC~5128, at least in the range of radial distances probed. The detected median metallicity gradient within the ACS field at $\sim 40$\,kpc from the centre of the galaxy should thus not reflect a global metallicity gradient, which would give unreasonably high metallicities if extrapolated inward. The median stellar metallicity gradient is most likely related to localised chemical variations that have not yet been fully blended into the diffuse smooth halo. The spatial correspondence and the similar metallicities between the small-scale chemical variations and the extended globular cluster suggest that this object is likely to be related to the same event that has led to the formation of the localised chemical variations. \section{Summary} \label{summary} In this paper, we have presented a discovery and discussion of two new globular cluster candidates in the outer region ($d \sim 40\,$kpc from the centre) of the nearest giant elliptical NGC~5128 using very deep HST/ACS imagery. The first and brightest of these two objects is selected based on both the presence of a clear RGB sequence and its structural parameters. The second one was selected solely on basis of its morphological appearance, colour and luminosity, and its structural parameters. The properties of the bright cluster are consistent with its identification as an old, intermediate-metallicity globular clusters resembling in every way we can verify the faint and extended clusters populating the Milky Way outer halo. This is the first time a counterpart of this populations of Galactic globular clusters has been reported for any other galaxy. The combined properties of the extended clusters and the diffuse stellar halo are consistent with the view that the reported clusters were once associated with dwarf galaxies that have disrupted. Finally, our present work indicates that a search for faint, extended globular clusters in NGC 5128 is entirely feasible. By carefully searching just one deep ACS/WFC imaging field -- which in this context is essentially a random pointing in the halo -- we have successfully identified two very probable GC candidates and characterized their structural parameters. Extrapolating from just this one field of area 11.5 arcmin$^2$ to the entire NGC 5128 halo area would be extremely risky, but it is worth noting that the Milky Way has more than a dozen known GCs fainter than $M_V \simeq -5$ \citep{harris96}. NGC 5128 is a more luminous galaxy by an order of magnitude, so it seems quite likely that its vast outer halo should hold some hundreds of faint GCs awaiting discovery, along with their attendant substructures. \section*{Acknowledgments} MM thanks Gary Da Costa for providing some of the data shown in Fig. 8 in an electronic form. This research has made use of the SIMBAD database, operated at CDS, Strasbourg, France.
\section*{Introduction} Motivated by the recent progresses in ultracold alkali atoms in disordered traps \cite{Aspect08,Fallani08,Fortagh07}, Nattermann and Pokrovsky proposed \cite{Nattermann08} a semiquantitative approach for Bose-Einstein condensates in strong random potentials where standard perturbative methods are expected to fail. Their analysis is based on the evaluation of the mean-field energy due to the fluctuations of the random potential according to a scaling argument introduced by Larkin~\cite{Larkin70} for studying the effect of defects in flux line lattices. The scaling approximation requires that the correlation length of the disorder is much shorter than the correlation length of the liquid. In this paper, we implement the method of Nattermann and Pokrovsky in the framework of the time dependent variational method of Perez-Garcia {\it et al.}~\cite{Zoller97}. As a result, we obtain a simple set of equations, which describe the equilibrium and the low-energy dynamics at zero temperature of a Bose-Einstein condensate in a strong random potential. Beside the oscillator and the scattering length, a new quantity enters into the problem, namely the Larkin length ${\mathcal L}$ associated with the collective pinning of the condensate. The time-dependent variational method represents a very good approximation of the Gross-Pitaevskii equation \cite{Kimura02,Salasnich04}. Moreover, it can interpolate quite successfully from the low density regime to the strong coupling Thomas-Fermi gas. The present extension to disordered traps, however, presents some restrictions. Since the fluctuating center of attraction of the disorder may not coincide with the center of the harmonic trap, the model is fully consistent only when either the harmonic trap or the disorder is responsible for the localization of the condensate. Moreover, the Gaussian variational ansatz out-rules from the outset the solutions describing the multi-fragmented state of the condensate~\cite{Nattermann08}. Despite these limitations, novel quantitative predictions can be obtained with regard to the disorder dominated regime. In this limit the condensate becomes non-superfluid and is characterized by a {\it generalized} correlation length~\cite{heal} larger than the size of the cloud. The paper is organized as follows. In Section \ref{sec1a} we introduce the general formulation of the problem. We determine the size of the ground state at equilibrium and we calculate the expressions for the frequencies of the low-lying excitations. In Section \ref{sec1b} we discuss in detail the $3$D gas in presence of a strong delta-correlated disorder. We show that, reducing the atomic scattering length by means of Feshbach resonance techniques, the predictions of the theory for the single connected fragment might be tested experimentally. In Section \ref{sec1c} low-dimensional gases are considered. Section \ref{sec1d} contains some considerations on the applicability of the method in the case of disorder with finite correlation length. In Section \ref{sec2a} we introduce a long-range anisotropic dipolar interaction and we study the stability of the single connected fragment. The nature of fragmentation in presence of dipolar interaction is analyzed in Section \ref{sec2b}. The static on-average characteristic properties of the fragments are determined at the level of the Imry-Ma level of approximation~\cite{Imry75}. \section{Time dependent variational method with disorder} \subsection{Variational equations} \label{sec1a} We begin considering the Hamiltonian of a gas of trapped bosons interacting through $s$-wave scattering \begin{eqnarray} \label{eq:hamiltonian} \hat{H}=\int d^3x \hat\Psi^{\dagger}\left( -\frac{\hbar^2}{2m}\nabla^2+V_{trap}({\bf x})+U\left({\bf x}\right)+\frac{4\pi\hbar^2 a }{m}\hat\Psi^{\dagger}\hat\Psi \right)\hat\Psi, \end{eqnarray} where $a$ is the $s$-wave scattering length and the harmonic trapping potential is $V_{trap}({\bf x})=\sum_{i=x,y,z}m\omega_i^2 R_i^2/2$. The disorder potential $U({\bf x})$ is chosen to be Gaussian distributed characterized by the average values $\langle U\rangle=0$ and $\langle U({\bf x} )U({{\bf x}}')\rangle=\left(\kappa^2/b^3\right) K_{b}\left({\bf x}-{\bf x}'\right)$, where $b$ denotes the correlation length of the disorder. We will assume that $K_{b}({\bf x})$ is a smeared out $\delta$-function $K_{b}({\bf x})={\rm e}^{-{\bf x}^2/b^2} /{(2\pi)^{3/2}}$. The variational calculation follows the same outline of the clean case. Therefore, in what follows, we refer the reader to~\cite{Zoller97} for the details of the derivation. We consider the normalized variational wave-function in the case of a $3$-D fully anisotropic configuration \begin{eqnarray} \label{eq:Zoller_Ansatz} \psi\left({\bf x},t\right)= \frac{N^{1/2}}{\pi^{3/4} {\tilde{R}^{\frac{3}{2}}(t)}} {e}^{-\sum_i^{3}\left[\frac{1}{2 {R_i^2(t)}}+i B_i(t)\right]x_i^2}, \end{eqnarray} where we have defined the geometric average $\tilde{R}=\left[ R_x(t) R_y(t) R_z(t)\right]^{1/3}$ and the variables $R_i$ and $B_i$ are time dependent variational parameters. {\color{black}The $R_i$'s are related to size of the system while the imaginary width $i B_i$ of the Gaussian ansatz is necessary to include the dynamics into the variational procedure. Without this latter the minimum principle would just lead to the condition of static equilibrium. } Physically, the true ground state of the stationary Gross-Pitaevskii equation in the presence of interactions and disorder is expected to deviate strongly from the ground state of the noninteracting system. Nevertheless, for a clean system, it has been shown that the dynamical trial wave-function~(\ref{eq:Zoller_Ansatz}) is a very good approximation of the Gross-Pitaevskii at finite $N$ \cite{Zoller97,Kimura02,Salasnich04}. Inserting the trial-wave function in the Lagrangian relative to the Hamiltonian of~(\ref{eq:hamiltonian}) we find \begin{eqnarray} \label{eq:Var_Lagr} L=L_{0}+\frac{\kappa}{\pi^{3/4}}\frac{N}{{\prod}_{i=1}^3 \left[2 R_i^2(t) +b^2\right]^{1/4}} \end{eqnarray} where $L_{0}$ is the well known contribution in absence of disorder~\cite{Zoller97}. The new term represents the on-average contribution of the disorder fluctuations according to the scaling argument of Nattermann and Pokrovsky~\cite{Nattermann08,Larkin70}. It introduces a new relevant length scale, namely the Larkin length ${\mathcal L}=({\color{black}2^{3/2}}\pi^{3/2}\hbar^4/m^2\kappa^2)$, associated with the {\color{black} pinning energy due to the disorder}~\cite{meanfreepath}. {\color{black} The length scale ${\mathcal L}$ was introduced by Larkin~\cite{Larkin70} in connection with the onset of collective pinning of vortex lines in type-II superconductors. In the case of a trapped gas, the analogous of the (delocalizing) elastic energy relative to the lattice distortion corresponds to the kinetic energy of the atoms. } Since the scaling argument requires to have {\color{black}both ${\cal L}$ and $b$ much smaller than the generalized healing length $\xi_{\rm heal}$~\cite{heal}, {\it i.e.}} ${\mathcal L}\ll \xi_{\rm heal}{\color{black} }$ and $b\ll \xi_{\rm heal}$, the validity of ~(\ref{eq:Var_Lagr}) is in general restricted to the quantum limit $b\leq {\mathcal L}$. The classical limit $b\gg {\mathcal L}$, when many levels occupy the typical well, will be briefly discussed separately. In what follows it is convenient to rescale the quantities in unities of the harmonic oscillator such as $r_i=R_i/{\ell}$ and $({\tilde b}/r_i)\equiv(b/\ell)(\ell/R_i)$ where $\ell=\sqrt{\hbar/m \omega}$ is the harmonic oscillator length. The anisotropy of the external trapping is taken into account by setting $\omega_i=\lambda_i \omega$. Without loss of generality we will consider a trap with cylindric symmetry with anisotropy factors $\lambda_x=\lambda_y=1$ and $\lambda_z=\lambda$. In absence of disorder this implies that the angular momentum along the $z$ axis is conserved and we can label the modes by the azimuthal angular quantum numbers $m$. {\color{black} Next, we derive the Euler-Lagrange equations for the variables $R_i(t)$ and $B_i(t)$ relative to the Lagrangian~(\ref{eq:Var_Lagr}). Then, eliminating the $B_i$'s variables, we obtain a closed system of differential equations for the $R_i$'s. These equations can be viewed as describing the motion of a point-like particle in an effective potential. The position of equilibrium at rest of the particle is determined by the minima of the effective potential. Using rescaled unities $r_{0x}=r_{0y}$ and $r_{0z}$, this leads ultimately to the conditions } \begin{eqnarray} \label{eq:equil_cond_x} r_{0x}^4+\gamma \frac{r_{0x}}{\sqrt{r_{0z}}} {f_{x}} {\prod_{j=1}^3 f_{j}^{1/4}} =1 + \alpha\frac{1}{r_{0z}}\\ \label{eq:equil_cond_z} \lambda^2 r_{0z}^4+ \gamma\frac{r_{0z}^{3/2}}{r_{0x}} {f_{z}} {\prod_{j=1}^3 f_{j}^{1/4}} =1+ \alpha\frac{ ~r_{0z}}{r_{0x}^2}, \end{eqnarray} where we have defined the functions {\small $f_{i}\left(r_{0i}\right)\equiv {\color{black}2}\left[2+\left({\tilde b}/{r_{0i}}\right)^2\right]^{-1}$} and {\small$g_{i}\left(r_{0i}\right)\equiv \left({{\tilde b}}/{r_{0i}}\right)^2 {f_{i}}/{\color{black}2}$, } and we have introduced the Thomas-Fermi parameter $\alpha=\sqrt{{2}/{\pi}}\left({N a}/{\ell}\right)$ together with the disorder strength parameter $\gamma=\sqrt{{\ell}/{\mathcal L}}$. In these two latter equations, the l.h.s. describes the confinement due to the trap and to the disorder. In the r.h.s. this effect is counterbalanced by the kinetic energy and the repulsive mean-field interactions. The limit $\mathcal L\rightarrow \infty$ corresponds to the theory for a clean system~\cite{Zoller97}. {\color{black}The low-lying excitations of the system can be calculated by means of a harmonic expansion $r_{i}(t)=r_{0i}+\delta r_i(t)$ around the equilibrium position~(\ref{eq:equil_cond_x}) and (\ref{eq:equil_cond_z}). Making a dynamical ansatz for the $\delta r_i(t)$, the calculus of the frequencies for the small oscillations is reduced to the calculation of the eigenvalues of a $3\times 3$ real symmetric matrix ${\rm A}_{ij}$} whose elements are \begin{eqnarray} \label{eq:matrx_el_gen} {\rm A}_{11}=&{\rm A}_{22}=1+\frac{3}{r_{0x}^4}+\frac{2\alpha}{{r_{0x}^4 r_{0z}}}- \frac{3\gamma}{2{r_{0x}^3 \sqrt{r_{0z}}}} f^{3/2}_{x}f^{1/4}_{z}\left\{ 1-\frac{5}{3}g_{x}\right\} \nonumber\\ {\rm A}_{12}=&\frac{\alpha}{{r_{0x}^4 r_{0z}}}- \frac{\gamma}{2{r_{0x}^3 \sqrt{r_{0z}}}} f^{3/2}_{x}f^{1/4}_{z}\left\{ 1-g_{x}\right\} \nonumber\\ {\rm A}_{13}=&\frac{\alpha}{{r_{0x}^3 r_{0z}^2}} -\frac{\gamma}{2{r_{0x}^2 {r_{0z}}^{3/2}}} f^{3/2}_{x}f^{1/4}_{z}\left\{ 1-g_{z}\right\} \nonumber\\ {\rm A}_{33}=&\lambda^2+\frac{3}{r_{0z}^4}+ \frac{2\alpha}{{r_{0x}^2 r_{0z}^3}} - \frac{3\gamma}{2{r_{0x} {r_{0z}^{5/2}}}} f^{1/2}_{x}f^{5/4}_{z}\left\{ 1-\frac{5}{3}g_{z}\right\}. \end{eqnarray} We denote the eigenvalues as $\omega_{\rm a,b,c}$. In absence of disorder $\omega_{\rm a,c}$ corresponds to the frequencies of the quadrupole modes with quantum numbers $n=0$ and $l=2$, while $\omega_{\rm b}$ is the frequency of the monopole mode with $n=1$, $l=0$. It is important to remember that the scaling argument leading to~(\ref{eq:Var_Lagr}) applies under the strong disorder condition $\ell \gg{\mathcal L}$. Since the center of the attractive domain formed by the disorder may not coincide with the center of the harmonic trap, the variational ground state is exact only when the system is trapped either by the static fluctuations of the random potential or by the harmonic trap. {\color{black} In principle, one could try to extend the ansatz~(\ref{eq:Zoller_Ansatz}) including an additional variational parameter which describes the center of the condensate. However, the disorder term in~(\ref{eq:Var_Lagr}) would not bring any new contributions to the equation of motion of the new variable with respect to that of the clean model~\cite{Zoller97}. This shows that the interplay between the two different confining mechanisms cannot be incorporated in the theory, which is a direct consequence of the {\it on-average} nature of the Larkin energy considered in~(\ref{eq:Var_Lagr})}. Therefore, the crossover region has to be considered {\color{black} at best} as an extrapolation. Moreover, our ansatz for the wave-function cannot describe a multi-fragmented state, which is expected to occur in a wide range of parameters~\cite{Nattermann08,frgm}. In this case our solution may occur as a metastable state \cite{Nattermann08}. Due to these limitations, in the next sections we will focus our analysis mainly to the single fragment non-superfluid regime where the cloud is trapped by the disorder and the physics is dominated by the Larkin length. \begin{figure} \vspace{2cm} \includegraphics[width=145mm,height=40mm]{figura1_a_b.eps} \caption{ (a) Frequency of the breathing mode $\omega_{\rm b}$ (red solid line) and of the quadrupole ${\rm c}$-mode (blue solid line) as a function of the particle number. The upper horizontal green dashed lines describe the analytic results of~(\ref{eq:mode_b_strg_dis_b_0_smN})- (\ref{eq:mode_c_strg_dis_b_0_smN}) while the lower dashed magenta correspond to the Thomas-Fermi limit $\omega_b=\sqrt{5}\omega$ and $\omega_{\rm c}=\sqrt{2}\omega$. We have considered a sample of $^{87}$Rb atoms confined in an isotropic trap with frequency $\omega=2\pi 50$ Hz and with scattering length $a=50 a_0$, where $a_0$ is the atomic Bohr radius. We have considered delta correlated disorder with ${\mathcal L}=200 a$ which amounts to $\ell/{\mathcal L}\approx 2.9$. (b) Same frequencies as in (a) and the ratio $\omega_b/\omega_c$ (magenta line), for a very shallow trap with $\omega=2\pi 10$ Hz and for a very small value of the scattering length $a=10a_0$. The Larkin length is $\mathcal{L}= 2000a$ which means $\ell/\mathcal{L}\approx 3.2$. } \label{figureA} \end{figure} \subsection{Strong disorder with zero correlation length} \label{sec1b} In the limit of very short correlated disorder $b\leq{\mathcal L}$ the corrections due to finite correlation length are very small. Therefore, as first approximation, we can consider a delta correlated disorder, and put $b=0$, {\color{black}$f_i=1$} and $g_i=0$ in~(\ref{eq:Var_Lagr})-(\ref{eq:matrx_el_gen}). {\color{black} Moreover, in this limit and for low energy oscillations, the wave-length of the modes is much larger than the distance over which the disorder varies. Therefore, the oscillations can be considered self-averaging and we expect the values of the frequencies to be independent of the specific realization.} At low densities and strong disorder, when both the radial and the axial oscillator lengths of the traps are larger than the Larkin length, the gas is confined mainly by the random potential and the cloud is spherically symmetric. The equilibrium conditions in~(\ref{eq:equil_cond_x}) and ~(\ref{eq:equil_cond_z}), which determines the size of the system, can be approximated by $\gamma\sqrt{r_0}\simeq1+\alpha/r_0$. \noindent In the limit $N\ll ({\mathcal L/a})$ the interactions can be neglected, the size of the cloud is $R\simeq {\mathcal L}$ \cite{mettinota} and we find \begin{eqnarray} \label{eq:mode_b_strg_dis_b_0_smN} \omega_{a}=\omega_b=&c_0\omega\left({\ell}/{\mathcal L}\right)^2\\ \label{eq:mode_c_strg_dis_b_0_smN} \omega_{c}=\left(c_0/2\right)&\omega\left({\ell}/{\mathcal L}\right)^2, \end{eqnarray} where the constant $c_0$ depends on the normalization of the variational trial wave-function we have chosen. Nevertheless, although the exact physical value of the constant $c_0$ cannot be rigorously determined by our approach, the ratio $\omega_{b}/\omega_{c}=2$ of the quadrupole and the monopole does not depend on it. For larger number of particles ${\mathcal L}\ll N a\ll \left({\ell}/{\mathcal L} \right)^{5/7}\ell$ the size of the system is determined by the competition between disorder and interaction and we find \begin{eqnarray} \label{eq:mode_a_strg_dis_b_0_TF} &\omega_a=\omega_c\propto \omega \frac{\ell^2}{(Na)^{7/6}{\mathcal L}^{5/6}} \left(\frac{{\mathcal L}}{Na}\right)^{1/6} \\ \label{eq:mode_b_strg_dis_b_0_TF} &\omega_b\propto \omega \frac{\ell^2}{(Na)^{7/6}{\mathcal L}^{5/6}}. \end{eqnarray} In this interval the generalized healing length $\xi_{\rm}$ reaches at some point the size of the system which is of the order of $(N a)^{2/3}{\mathcal L}^{1/3}$. Moreover, the two curves exhibit an ``avoided crossing-like" feature. For even larger number of particles the external harmonic potential dominates over the disorder and the frequencies approach the well-known Thomas-Fermi \cite{Stringari96}. The crossover between the different regimes of the values of the quadrupole and monopole modes is illustrated in figure~\ref{figureA}{a} for some typical experimental parameters. At $N\gg \left({\ell}/{\mathcal L}\right)^{5/7}(\ell/a)$ the crossover to the Thomas-Fermi regime of the clean theory occurs~\cite{Nattermann08}. Deep into the Thomas-Fermi regime the system is superfluid with $\xi_{\rm heal}\ll {\mathcal L}$. {\color{black}The critical number of particles $N_{\rm c}$ where transition to superfluid occurs can be estimated by equating the Larkin length with the superfluid healing length obtained from eqs.~(\ref{eq:equil_cond_x})-(\ref{eq:equil_cond_z}) and the definition given in~\cite{heal}. This leads to $N_{\rm c}\sim\ell^6/a {\cal L}^5$ in agreement with the theory of~\cite{gian09,gian09b}.} In the Thomas-Fermi limit the disorder can be treated perturbatively and a small negative linear shift in $\xi_{\rm heal}/{\mathcal L}$ would be expected~\cite{Falco07}. However, in figure~\ref{figureA}{a} a positive shift of the frequency of the breathing mode is found. This is not surprising since the present theory has to be considered as exact only in the opposite limit $\xi_{\rm heal}\gg {\mathcal L}$ where disorder is dominating. The frequencies plotted in figure~\ref{figureA}{a} refer to a single connected condensate. However, above $N\geq{\mathcal L}/a$ the single connected condensate is a higher energy metastable state, since the ground state is expected to undergo fragmentation~\cite{Nattermann08}. In the fragmented state, we do not expect sharply defined frequencies. More likely, they should be distributed in some interval. {\color{black} Note that the possibility to observe the single connected condensate, in the regime where fragmentation is expected, could arise by means of a sudden decrease of the oscillator frequency. However, it is difficult to make further estimates about the different timescales involved in the experiment. In order to prevent fragmentation,} we would like to push the single fragment solution at larger particle numbers. For a given disorder (${\mathcal L}$) and trap configuration ($\ell$) such that strong disorder condition $\ell\gg {\mathcal L}$ is satisfied, this can be achieved by maximizing the ratio ${\mathcal L}/a$. Experimentally, this can be obtained by tuning the scattering length close to zero ({\it zero crossing}) by means of a Feshbach resonance as proposed in Fig.~\ref{figureA}{b}. In the case of negative scattering length there is only a metastable state of finite radius. This state becomes unstable at a critical particle number $N_c$. In Fig.~\ref{figure2}a, we plot the frequencies of the three modes in the case of an attractive interaction for an isotropic trap with $\omega=2\pi 25$ Hz, $a=-100 a_0$ and ${\mathcal L}=200 a$. The divergence signals the instability of the gas when the number of particles reaches the critical value $N_c$. \begin{figure} \vspace{1cm} \includegraphics[width=145mm,height=40mm]{figura2_ab.eps} \caption{(Colour online)(a) Example of the oscillations frequencies as function of the number of particles for negative scattering length. The upper solid (green) line describes the $a,b-$modes while the lower (red) shows the $c-$mode. The dashed (yellow) lines indicates the asymptotic solutions~(\ref{eq:mode_b_strg_dis_b_0_smN})-(\ref{eq:mode_c_strg_dis_b_0_smN}). (b) Frequency of ``quadrupole" mode $c$ (red solid line). The upper dot-dashed light blue line corresponds to the analytic result of~(\ref{eq:mode_c_strg_dis_b_0_smN_oneD}). The solid dark-blue line indicates the frequency $\sqrt{3}~\omega_{\parallel}$ characteristic of the $1D$ mean-field theory, while the solid light-blue represents the $\lambda\rightarrow 0$ limit of the $3D$ Thomas-Fermi frequency $\sqrt{5/2}~\omega_{\parallel}$. The clean theory is described by the dashed green line which at low density approaches the free h.o. value $2\omega_{\parallel}$ (magenta dot-dashed). The inset shows the details of the crossover from the $1D$ mean-field regime to the Thomas-Fermi condensate. } \label{figure2} \end{figure} \subsection{Anisotropic traps and lower dimensions} \label{sec1c} We have seen that, at low densities, when the physics is determined only by the balance between the random potential and the kinetic energy, the frequencies of the oscillations do not depend on the external harmonic potential. Therefore, the anisotropy of the trap becomes irrelevant. Increasing the number of particles, the mean-field interaction becomes important and the harmonic trap affects indirectly the results through the Thomas-Fermi parameter $\alpha$. At even larger particle number the system enters the Thomas-Fermi regime where the harmonic trap dominates over the disorder and the anisotropy plays its usual role \cite{Stringari96}. If we consider, for example, an elongated cigar along the axial $z$-direction, with $\omega_{x,y}\equiv\omega_{\perp}\gg \omega_{\parallel}\equiv\omega_z$, this description holds as long as we have ${\mathcal L}\ll \ell_{\perp}\ll \ell_{\parallel}$. However, when $\ell_{\perp}\ll {\mathcal L}\ll \ell_{\parallel}$ the system can pass through the different regimes in the axial directions while its ground state in the radial direction is frozen to that of the harmonic oscillator of the radial harmonic confinement. Herewith, we limit our discussion to the single connected solution, neglecting the fragmented low-dimensional condensate where it occurs~\cite{Nattermann08}. Using the definitions $a_{1D}=\ell_{\perp}^2/a$ and ${\mathcal L}_{1D}=({\mathcal L}\ell_{\perp}^2)^{1/3}$ we have that for small $N\ll a_{1D}/ {\mathcal L}_{1D}$ the axial radius is determined by the equation $\gamma r_{z0}^{3/2}=1$. This leads to the results \begin{eqnarray} \label{eq:mode_b_strg_dis_b_0_smN_oneD} &\omega_a=\omega_{b}=2\omega_{\perp}\\ \label{eq:mode_c_strg_dis_b_0_smN_oneD} &\omega_{c}\propto \omega_{\perp}\left({\ell_{\perp}}/{\mathcal L}\right)^{2/3}. \end{eqnarray} At larger $N$ such $ a_{1D}/ {\mathcal L}_{1D} \ll N \ll a_{1D}/{\mathcal L}_{1D}(\ell_{\parallel}/{\mathcal L}_{1D})^{4/5}$ the axial radius is determined in the leading order by the equation $\gamma r_{z0}^{3/2}=\alpha r_{z0}$ and we have in first approximation \begin{eqnarray} \label{eq:mode_a_strg_dis_b_0_TF_oneD} &\omega_a=\omega_b=2\omega_{\perp} \\ \label{eq:mode_c_strg_dis_b_0_TF_oneD} &\omega_{c}\propto \omega_{\perp}\left({\ell_{\perp}}/{\mathcal L}\right)^{3/2} \left({\ell_{\perp}}/{N a}\right)^{5/2}. \end{eqnarray} At even larger values the Thomas-Fermi regime is realized in the axial direction while in the radial direction the ground state wave-function can still be determined by the free harmonic oscillator. Therefore, we have that the frequency of the breathing mode approaches the value $\omega_c= \sqrt{3}\omega_{\parallel}$ characteristic of the so-called $1D$ mean-field regime \cite{Stringari97}. In order to appear, this regime requires the two conditions $N a\lambda/\ell_{\perp}\ll 1 $ and $(N a/\sqrt{\lambda}\ell_{\perp})^{1/3}\gg 1$ \cite{Menotti02}. Finally, for $N a\lambda/\ell_{\perp}\gg 1$ the gas enters the full Thomas-Fermi regime and the "quadrupole" mode approaches the strongly anisotropic $3D$ Thomas-Fermi result \cite{Stringari96} $\omega_c=\sqrt{5/2}\omega_{\parallel}$ while $\omega_a=\sqrt{2}\omega_{\perp}$ and $\omega_b=2\omega_{\perp}$. The full crossover is shown in Fig.~\ref{figure2}b for the lowest axial mode $c$, in a $^{87}$Rb condensate confined in a strongly anisotropic cigar-shaped trap with $\omega_{\perp}=2\pi 150 $Hz, $\omega_{\parallel}=0.01 \omega_{\perp}$, $a=150 a_0$, and ${\mathcal L}=300 a$. For these parameters, $\ell_{\perp}/{\mathcal L}\approx 0.36$, and $\ell_{\parallel}/{\mathcal L}\approx 3.7$. Moreover, the gas cannot enter the strong interacting Tonks regime since it would require to fulfill the two conditions $\ell_{\perp}/a< 10^2$ and $N\lambda< 1$ simultaneously~\cite{Menotti02}. In our case, we have $\sqrt{\lambda}\ell_{\perp}/a\approx 11$. Nevertheless, the Tonks gas remains beyond the reach of our mean-field approach. The treatment of a $2$D disc-like geometry trap when $\lambda\rightarrow\infty$ follows essentially the same outline. In this case we define $\omega_{x,y}\equiv\omega_{\parallel}\ll \omega_{\perp}=\omega_z$. Similar to the $1$D configuration, a non trivial interplay between disorder and low-dimensionality arises only when $\ell_{\perp}\ll {\mathcal L}\ll \ell_{\parallel}$. Except that for very large $N$, the ground state in $z$-radial direction coincides with the lowest eigenstate of the harmonic trap, namely $r_{0z}^4=1/\lambda^2$. The $2D$ analog of ~(\ref{eq:mode_c_strg_dis_b_0_smN_oneD}) and~(\ref{eq:mode_c_strg_dis_b_0_TF_oneD}) are respectively $\omega_c\propto \omega_{\perp}{l_{\perp}}/{\mathcal L}$ in the small $N$ limit and $\omega_c\propto({l_{\perp}}/{\mathcal L})({l_{\perp}}/{N a})^{3/2}$ in the moderate interacting regime. At larger values of $N$, for sufficiently strong anisotropies, the size of the cloud in the $x,y$ plane is determined by the competition between harmonic trap and interactions while in the $z$-direction the ground state can still be that of the free harmonic oscillator. This is the $2D$ analog of the $1D$ mean-field regime for which $\omega_c=2 \omega_{\parallel}$ \cite{Ma00}. Finally when $N\rightarrow \infty$ the $\lambda\rightarrow\infty$ limit of the $3D$ Thomas-Fermi is recovered where $\omega_c=\sqrt{10/3}~\omega_{\parallel}$ \cite{Stringari96}. \subsection{Finite correlation length} \label{sec1d} The delta correlated disorder approximation discussed so far, is appropriate for disorder with very short correlation length, {\it i.e.} when $b\leq {\mathcal{L}}$. In this case, the correlation length of the disorder is generally much shorter than the correlation length of the system, which represents a necessary condition for applying the Larkin scaling argument. The small corrections due to finite $b$ can be taken into account using the full equations~(\ref{eq:equil_cond_x})-(\ref{eq:matrx_el_gen}). Differently, when $b \gg {\mathcal{L}}$, the use of the expression for disorder energy in~(\ref{eq:Var_Lagr}) is somehow questionable. {\color{black} Moreover, also self-averaging is expected to be much less efficient than in the limit delta correlated disorder.} Therefore, we conclude that the description based on~(\ref{eq:equil_cond_x})-(\ref{eq:matrx_el_gen}) holds as far as the limiting case of delta correlated disorder is approached. Nevertheless, it is interesting to make the following remark. The classical limit $b \gg {\mathcal{L}}$ has been recently investigated in~\cite{gian09,Shklovskii08}. Near one of its typical minima, the random potential can be approximated by a harmonic potential well of depth $U_0=\kappa/b^{3/2}$ and width $b$. The bound state is located very close to the minima of the random potential and the number of levels in a well is large. From $U_0\equiv m \omega_b b^2/2$ we obtain the oscillator length of the typical fluctuating well as $\ell_b=b\left({\mathcal{L}}/b\right)^{1/8}$~\cite{gian09b,Shklovskii08}. Interestingly, this length can be obtained from~(\ref{eq:equil_cond_x}) in the limit of small number of particle where the harmonic trap and the mean-field energy can be neglected. At larger particle number the size of the fragment is determined by the competition of the mean-field interaction and the disorder energy. This yields $R\sim(Na{\mathcal{L}}^{1/2} b^{7/2})^{1/5}$, which agrees with the size of the fragments in the multi-domain state found in~\cite{gian09b} by different methods. This analysis reveals that the the Gaussian ansatz~(\ref{eq:Zoller_Ansatz}) is able to reproduce the ground-state localized state also for disorder with long-range correlated potential. \begin{figure} \vspace{1cm} \includegraphics[width=150mm,height=40mm]{figura3_a_b.eps} \caption{(Colour online)(a) The size of the single fragment of $^{52}$Cr atoms along the radial (green) and the axial (magenta) directions as a function of the ``relative" dipolar strength. The parameters are $N=1000$, $a=10 a_0$, $\lambda=1$, $\omega=2 \pi 10$ Hz, ${\mathcal L}\sim 2000 a$, and $\ell/{\mathcal L}\sim3.2$. The inset shows their ratio $r_{0x}/r_{0z}$. (b) Dependence of critical value of the dipole interaction ($a_{\rm dip}/a$) from the particle number for ${\mathcal L}\sim 2000 a$ (red) and for ${\mathcal L}\sim 1000 a$ (green). The other parameters are taken as in figure~\ref{figure3}(a). The dashed line at $N=150$ is just a guide for the eye. } \label{figure3} \end{figure} \section{Dipolar gas} \subsection{Stability of the gas} \label{sec2a} The time dependent variational approach can be extended to the presence of the long range anisotropic dipolar interaction~\cite{Yi01}. If we assume the dipoles aligned along the $z$-direction, the interaction between two polarized dipoles can be written as \begin{equation} V\left(\bf r\right) =\frac{\mu_0\mu^2}{4\pi}4\sqrt{\frac{\pi}{5}} \frac{Y_{20}(\theta)}{r^3}, \label{eq:dippot} \end{equation} where $\theta$ is the angle between ${\bf r}$ and the direction along which the dipoles are pointing. The constant $\mu_0$ is the magnetic constant and $\mu$ is the magnetic moment of the atoms. The strength of the dipole interaction relative to the short range potential will be expressed through the dimensionless quantity $\varepsilon =\sqrt{{\pi}/{5}}\left(\mu_0\mu^2 m/4 \pi^2\hbar^2 a\right)$. Alternatively, the strength of the two potentials can be compared by defining a characteristic length scale associate to the dipole-dipole interaction. From the uncertainty principle, we can define $a_{\rm dip}$ as the distance at which the dipolar potential energy~(\ref{eq:dippot}) equals the kinetic energy. This yields \begin{equation} a_{\rm dip} =\sqrt{\frac{\pi}{5}}\frac{\mu_0\mu^2 m}{4 \pi^2 \hbar^2 }, \end{equation} where we have chosen the prefactor such that $\varepsilon=a_{\rm dip}/a$ \cite{footnoteA}. In presence of disorder and dipolar interaction, the equilibrium conditions of~(\ref{eq:equil_cond_x}-\ref{eq:equil_cond_z}) become \begin{eqnarray} \label{eq:equil_cond_x_dip} r_{0x}^4+\gamma \frac{r_{0x}}{\sqrt{r_{0z}}} {f_{r_{0x}}} {\prod_{j=1}^3 f_{r_{0j}}^{1/4}} =1 + \alpha\frac{1}{r_{0z}}\left[1-\varepsilon {\mathcal F}\left(\frac{r_{0x}}{r_{0z}}\right) \right]\\ \label{eq:equil_cond_z_dip} \lambda^2 r_{0z}^4+ \gamma\frac{r_{0z}^{3/2}}{r_{0x}} {f_{r_{0z}}} {\prod_{j=1}^3 f_{r_{0j}}^{1/4}} =1+ \alpha\frac{ ~r_{0z}}{r_{0x}^2}\left[1-\varepsilon {\mathcal G}\left(\frac{r_{0x}}{r_{0z}}\right)\right]. \end{eqnarray} with \cite{Yi01} ${\mathcal F}\left(\xi\right)=\left[\sqrt{5 \pi}/6(1-\xi^2)^2\right]\left[ -4\xi^4-7\xi^2+2+9\xi^4{{\mathcal H}\left(\xi\right)}\right]$, and ${\mathcal G}\left(\xi\right)=\left[\sqrt{5 \pi}/3(1-\xi^2)^2\right]\left[ -2\xi^4+10\xi^2+1-9\xi^2{{\mathcal H}\left(\xi\right)}\right]$, and ${\mathcal H}\left(\xi\right)={\rm{Arctanh}{\sqrt{1-\xi^2}}}/{\sqrt{1-\xi^2}}$. \begin{figure} \vspace{1cm} \includegraphics[width=150mm,height=40mm]{figura4_a_b.eps} \caption{(Colour online)(a) Energy of a $^{52}$Cr condensate in a strong random potential in absence of dipolar interaction (red) and for $a_{\rm dip}=0.40 a$ (blue) as function of $N$. The other parameters are $a=100 a_0$, $\lambda=1$, $\omega=2 \pi 50$ Hz, ${\mathcal L}\sim 30 a$, and $\ell/{\mathcal L}\sim 9.5$. The inset shows the correspondent energy per particle. (b) Typical deformation of the fragments as function of the relative dipolar strength for two different values of the disorder ${\mathcal L}\sim 50 a$ (green), ${\mathcal L}\sim 30 a$ (red). The other parameters have been chosen as in figure~\ref{figure4}a. $^{52}$Cr atoms, $a=100 a_0$, $\lambda=1$, $\omega=2 \pi 50$ Hz, upper ${\mathcal L}\sim 50 a$, lower ${\mathcal L}\sim 30 a$.} \label{figure4} \end{figure} Analogously to the clean gas, these equations have at most only one stable solution. Here we are interested in the regime of strong disorder and moderate interactions where the gas is confined mainly by the random potential and is constituted by a single connected fragment. In this limit the anisotropy of the harmonic trap does not play any role and, in absence of dipolar interaction, the cloud would have spherical symmetry. However, turning on the dipolar potential the fragment tends to become more prolate along the direction of the dipoles orientation. The situation is illustrated in figure~\ref{figure3}a, where we have considered a very short scattering length in order to shift towards large $N$ the onset of multiple fragmentation which occurs above $N\geq {\mathcal L}/a$. Moreover, such a small value of the scattering length, permits to access the strong dipolar regime $a_{\rm dip}/a\geq 1$ even for gases made of atoms with small magnetic moments. This regime has been recently realized in experiments with $^{52}$Cr condensates near a Feshbach resonance \cite{Pfau05a,Pfau08_Nature}. For any fixed $N$, the fragment collapses at some critical value of $a_{\rm dip}/a$. In figure~\ref{figure3}b we plot this threshold as function of the number of particles at fixed $a$ for two different values of the Larkin length. For given $a$ and $N$ at shorter Larkin length the condensate is less stable against the attractive dipole interaction. Note that a shorter Larkin length amounts to a stronger center of attraction due to the fluctuations of the random potential and thus to a smaller volume occupied by the fragment. \subsection{Fragmented state} \label{sec2b} So far we have not discussed the possibility of a multi-fragmented state since the variational method is based on a single connected ground state. However, in~\cite{Nattermann08} it has been shown that for a large region of the parameters space, strong disorder favours fragmentation. It is therefore interesting to analyze the situation when a dipolar interaction is present as well. In the spirit of~\cite{Nattermann08}, we evaluate the Hamiltonian operator on the ground state~(\ref{eq:Zoller_Ansatz}). Assuming cylindrical symmetry and rescaling energies in unities of the harmonic trap energy $\hbar\omega$, the energy per particle of the system can be written as \begin{eqnarray} \frac{\mathcal{E}}{\hbar\omega}= \frac{1}{2 r^2} \left( 1 +\frac{\sigma^2}{2} \right)+ \frac{1}{2} {r^2}\left(1+\frac{\lambda^2}{2\sigma^2} \right) -\sqrt{\frac{\ell}{{\mathcal L}}}\sqrt{\frac{{\sigma}}{{r^3 }}} +\frac{\alpha}{2}\frac{\sigma}{r^3} \left[ 1-\varepsilon I_1\left(\sigma \right) \right]\nonumber, \end{eqnarray} in terms of the variational parameters $r\equiv r_{x}$ and $\sigma=r/r_z$. The {\it deformation} function $I_1(\xi)={\sqrt{5 \pi}}\left[ 1+2 \xi^2-3\xi^2{{\mathcal H}\left(\xi\right)} \right]/{3\left(1-\xi^2\right)}$ is continuous, monotonous and positive for $\xi<1$ and negative for $\xi>1$. In absence of dipolar interaction and for a strong given disorder potential the total energy $E(N)={N\mathcal E}$ has a minimum $E_1(N_1)$ at negative energy as shown in figure~\ref{figure4}a. According to the Imry-Ma criterium~\cite{Nattermann08}, at $N$ larger than $N_1$ it is energetically favourable to split the gas into fragments of energy $E_1$. Eventually, at even larger $N$, the oscillator energy per particle of the fragmented state reaches the energy per particle from the disorder and the crossover to the Thomas-Fermi regime occurs \cite{Nattermann08}. The effect of the dipolar interaction on the phenomenon of the fragmentation can be seen as follows. The attractive dipole interaction shrinks the total condensate volume causing an increase of the kinetic energy which tends to delocalize the condensate against the collective pinning induced by the disorder. This is consistent with the numerical results shown in figure~\ref{figure4}a. At fixed disorder, finite values of the dipolar interaction shift upward the minimum of the energy $E={N\mathcal E}$ until it becomes positive. Then, strictly speaking, the possibility of having fragmentation is ruled out at least at the level of the Imry-Ma argument. Nevertheless, at sufficiently small dipolar interaction fragmentation can occur. In the presence of anisotropic interactions the Imry-Ma \cite{Imry75} argument does not imply that domains have to be spherical. The typical fragment will be preferentially cigar-shaped with the long axis parallel to the dipoles \cite{Nattermann88}. This means that, while in a single realization of the disorder the various fragments should appear with random shape, they will have on average a characteristic deformation. The dependence on the dipolar interaction of the typical deformation can be estimated evaluating the equilibrium radii $r_{0x}$ ad $r_{0z}$ in the minimum of the energy curve $E(N)$. The result obtained is shown in figure~\ref{figure4}b for two different realization of the disorder. The curves stop where the Imry-Ma argument for the existence of fragmentation becomes invalid. Note that, for the chosen parameters, this boundary is smaller than the critical dipolar strength at which the cloud would collapse. \section*{Conclusions} We have proposed a hybrid approach, originated from the variational method for the Gross-Pitaevskii theory and the Larkin-Imry-Ma scaling argument of~\cite{Nattermann08}, aimed to study zero temperature Bose-Einstein condensates in strongly disordered traps. {\color{black} Similarly as for other disordered systems, the Larkin-Imry-Ma scaling analysis, despite its intrinsic limitations, allows to make semi-quantitative predictions in situations where no mean-field description seems possible~\cite{Hertz85}.} The theory addresses the problem of the condensate in the limit of strong disorder and moderate interaction. There, the dynamics of the condensate is dominated by a new length scale, namely the Larkin length, associated to the collective pinning of the condensate. {\color{black} In the case of isotropic interaction, we have investigated the stability and the low-lying excitations of the system at different dimensionalities when the ground state consists of a single connected condensate. We have suggested that, for given disorder strength and trapping parameters, this regime could be observed experimentally by using scattering length zero crossing \cite{Fattori08}. In the case of dipolar interactions, our theory offers in addition the possibility to investigate the effects of the anisotropic interaction on the phenomenon of fragmentation. Predictions about the typical deformation ratio of the fragments can be made. We conclude that,} the realization of types of disorder with correlation length smaller than all other length scales, could allow to test the theory in the laboratory. \ack{I would like to thank T. Nattermann and V. L. Pokrovsky for useful discussions.} \section*{Bibliography}
\section{Extended Irreversible Thermodynamics (EIT)} We begin with summarizing the basis of perfect fluid and classic laws of dissipations (e.g. Fourier and Navier-Stokes laws) in order to clarify the basis and need for EIT. The perfect fluid is a phenomenology assuming the \emph{local equilibrium} of fluid, which requires that each fluid element is in a thermal equilibrium state. By the local equilibrium assumption, any fluid element in perfect fluid evolves adiabatically and no entropy production arises in the fluid element. This contradicts the dissipative phenomena which are essentially the irreversible and entropy producing processes. Therefore, the basic equations of perfect fluid can not include any dissipation. Then we notice that \emph{the local equilibrium assumption is inconsistent with the irreversible nature of dissipative phenomena}. In other words, dissipations can not exist in local equilibrium systems. On the other hand, recall that the classic laws of dissipations are also the phenomenologies assuming the local equilibrium. Hence the classic laws lead inevitably an unphysical conclusion; an infinitely fast propagation of dissipations. While the infinitely fast propagation may be harmless to Newtonian theories, however, in relativistic theories, it gives rise to a serious problem; the violation of causality. The classic laws of dissipations can not be accepted as basic laws of relativistic dissipations. (See~\cite{ref:eit} for details of the violation of causality by the local equilibrium assumption.) From the above, it is recognized that we should abandon the local equilibrium assumption in order to obtain a consistent dissipative hydrodynamics. Therefore the idea of \emph{local \underline{non}-equilibrium} is necessary, which means that the fluid element is in a non-equilibrium state. A physically consistent phenomenology of dissipative hydrodynamics, which is based on the local \emph{non}-equilibrium idea, is already formulated. It is called the \emph{Extended Irreversible Thermodynamics} (EIT). As precisely explained in~\cite{ref:eit}, because of the local non-equilibrium idea, EIT describes the entropy production inside each fluid element, which results in a finite speed of propagation of dissipations in both non-relativistic and relativistic situations. The relativistic EIT is a causally consistent dissipative hydrodynamics~\cite{ref:israel,ref:is,ref:hl_causality,ref:hl_applicable}. Although the EIT is a dissipative ``hydrodynamics'', it is called ``thermodynamics''. This name puts emphasis on the replacement of local equilibrium idea with local \emph{non}-equilibrium one, which is a revolution in thermodynamic treatment of fluid element. The terminology ``EIT'' is found in~\cite{ref:eit} which is developed by experts in non-equilibrium physics. Contrary, the original works of relativistic dissipative hydrodynamics~\cite{ref:israel,ref:is,ref:hl_causality,ref:hl_applicable} put emphasis not on the thermodynamic revolution but on the preservation of causality. Here let us dare to use the term ``EIT'' since the local non-equilibrium nature of dissipative fluid is the physical origin of preservation of causality. Before showing the basic equations of EIT, we list the basic quantities; \begin{eqnarray*} u^{\mu}(x) &:& \mbox{velocity field of fluid (four-velocity of fluid element)}\\ \rho(x)\,\,\Bigl(\,=\frac{1}{V(x)}\,\Bigr) &:& \mbox{mass density ($\rho$) and specific volume (volume per unit mass, $V$)}\\ \varepsilon_{\rm ne}(x) &:& \mbox{non-equilibrium specific internal energy (internal energy per unit mass)}\\ p_{\rm ne}(x) &:& \mbox{non-equilibrium pressure}\\ T_{\rm ne}(x) &:& \mbox{non-equilibrium temperature}\\ q^{\mu}(x) &:& \mbox{heat flux vector}\\ \Pi(x) &:& \mbox{bulk viscosity}\\ \widetilde{\Pi}\,^{\mu\nu}(x) &:& \mbox{shear viscosity tensor}\\ g_{\mu\nu}(x) &:& \mbox{spacetime metric} \end{eqnarray*} Here $x$ denotes the dependence on spacetime point. All of the above quantities except for $u^{\mu}$ and $g_{\mu\nu}$ are the \emph{non-equilibrium thermodynamic state variables} which characterize the non-equilibrium state of each fluid element. The non-equilibrium state variables are classified into two categories: The quantities $\varepsilon_{\rm ne}$, $p_{\rm ne}$, $T_{\rm ne}$ and $\rho$ are the state variables which exit even at the local equilibrium limit, while the quantities $q^{\mu}$, $\Pi$ and $\widetilde{\Pi}\,^{\mu\nu}$ are the state variables which should vanish at the local equilibrium limit. We call the first category ($\varepsilon_{\rm ne}$, $p_{\rm ne}$, $T_{\rm ne}$, $\rho$) the \emph{non-equilibrium scalars}, and call the second category ($q^{\mu}$, $\Pi$, $\widetilde{\Pi}\,^{\mu\nu}$) the \emph{dissipative fluxes}. The dissipative fluxes are the origin of dissipative phenomena and make the fluid being non-equilibrium. The suffix ``ne'' for non-equilibrium scalars denotes ``non-equilibrium'', and the variables with this suffix have different value from an equilibrium case. Note that the definition of $\rho$ (or $V$) is the same for both equilibrium and non-equilibrium cases; to count the number of composite particles or measure the mass per unit volume. So the suffix ``ne'' is not given to $\rho$ and $V$. As the basic assumption of EIT, it is required that the dissipative fluxes are \emph{independent} non-equilibrium state variables. Furthermore, concerning the non-equilibrium scalars, it is also assumed that, as in the ordinary equilibrium thermodynamics, the number of independent non-equilibrium scalars is two for \emph{closed} systems which conserve the number of composite particles, and three for \emph{open} systems in which the number of composite particles changes. The remaining state variables are expressed as functions of the independent variables through the equations of state, e.g. the non-equilibrium specific entropy $s_{\rm ne}$ is a function of independent state variables $s_{\rm ne} = s_{\rm ne}(\varepsilon_{\rm ne},V,q^{\mu},\Pi,\widetilde{\Pi}\,^{\mu\nu})$, where $\varepsilon_{\rm ne}$ and $V$ are chosen as independent non-equilibrium scalars. Here note the fact that EIT, at least in its present status, is not necessarily applicable to any non-equilibrium state of dissipative fluid~\cite{ref:eit}. The consistent basic equations of EIT can be formulated for sufficiently weak dissipative fluxes, which means that the strength of non-equilibrium nature should not be so strong. For example, it is shown in~\cite{ref:hl_applicable} that the EIT preserves causality of heat flow if $\sqrt{q^\mu q_\mu}/\varepsilon_{\rm ne} \mathrel{\mathpalette\vereq<}}}{ 0.08$ for non-viscous fluid ($\Pi = 0$, $\widetilde{\Pi}\,^{\mu\nu} = 0$) under stationary and homogeneous (spatially one-dimensional) condition. We can recognize that the amount of energy transported by dissipative fluxes should be less than a few percent of the internal energy which includes mass energy. This efficiency of dissipative energy transfer is larger than the efficiency of hydrogen burning in a star ($\sim O(10^{-3})$ ). Therefore, although the EIT is applicable to a dissipative fluid whose local non-equilibrium states are not so far from local equilibrium states, it is expected that the EIT is applicable to many astrophysical systems. In the present status of EIT, the above restriction is reflected in the equations of state. The equations of state, e.g. $s_{\rm ne}(\varepsilon_{\rm ne},V,q^{\mu},\Pi,\widetilde{\Pi}\,^{\mu\nu})$, are expanded up to second order of dissipative fluxes about the \emph{fiducial equilibrium state}. Here the \emph{fiducial equilibrium state} is defined as an equilibrium state of fluid element of imaginary perfect fluid possessing the same fluid velocity $u^{\mu}(x)$ and mass density $\rho(x)$ with our actual dissipative fluid. Then, as explained in~\cite{ref:eit}, the evolution equations of dissipative fluxes are obtained under two requirements; (i)~positivity of entropy production rate at each fluid element, and (ii)~consistency with experimentally determined phenomenology of relaxation processes of dissipations. The resultant evolution equations are; \begin{eqnarray} \label{eq:heat} \tau_{\rm h}\,\cd{q}\,^{\mu} &=& - q^{\mu} - \lambda T\,\cd{u}\,^{\mu} + \tau_{\rm h}\,(q^{\nu}\cd{u}_{\nu})\,u^{\mu} - \lambda\,\Delta^{\mu\nu}\,\Bigl(\, T_{,\nu} - T^2\,\bigl(\,\beta_{\rm hb}\,\Pi_{,\nu} + \beta_{\rm hs}\,\widetilde{\Pi}\,^{\,\,\,\alpha}_{\nu\,\,\,\,;\alpha} \,\bigr) \Bigr) \\ \tau_{\rm b}\,\cd{\Pi} &=& - \Pi - \zeta\,u^{\mu}_{\,\,\,;\mu} + \beta_{\rm hb}\,\zeta\,T\,q^{\mu}_{\,\,\,;\mu} \\ \label{eq:shear} \tau_{\rm s}\,\bigl(\,\widetilde{\Pi}\,^{\mu\nu}\bigr)^{\cdot} &=& - \widetilde{\Pi}\,^{\mu\nu} + 2\,\tau_{\rm s}\,\cd{u}_{\alpha}\,\widetilde{\Pi}\,^{\alpha (\nu}\,u^{\nu)} - 2\,\eta\,\Bigl[\, u^{\mu;\nu} - T\,\beta_{\rm hs}\,q^{\mu;\nu} \,\Bigr]^{\circ} \,, \end{eqnarray} where $T$ is the temperature of fiducial equilibrium state. Here, definitions of mathematical symbols are; $\cd{Q} := u^{\mu}Q_{;\mu}$, $\Delta^{\mu\nu} := u^{\mu} u^{\nu} + g^{\mu\nu}$ and $[A_{\mu\nu}]^{\circ} := \Delta^{\mu\alpha}\Delta^{\nu\beta}A_{(\alpha\beta)} - (1/3)\Delta^{\mu\nu}\Delta^{\alpha\beta}A_{\alpha\beta}$. And the meanings of phenomenological coefficients are; $\tau$'s are relaxation times of dissipative fluxes, $\lambda$ is heat conductivity, $\zeta$ and $\eta$ are respectively bulk and shear viscous rates, and $\beta_{\rm hb}$ and $\beta_{\rm hs}$ are interaction coefficients between $q^{\mu}$ and $\Pi$ or $\widetilde{\Pi}\,^{\mu\nu}$. For example, $\beta_{\rm hs}$ means that the heat flow arises in a shear viscous flow, also shear viscosity arises in a flow with heating. Also $\beta_{\rm hb}$ means the same between $q^{\mu}$ and $\Pi$. The values of those phenomenological coefficients should be determined by some experiments. Furthermore note that, because dissipative phenomena are not time reversal, these evolution equations of dissipative fluxes are also not time reversal. In reference~\cite{ref:is} which is one of the original works of relativistic EIT, the \emph{molecular viscosity} are assumed as the physical origin of viscosities $\Pi$ and $\widetilde{\Pi}\,^{\mu\nu}$. However in Newtonian accretion disk theories~\cite{ref:disk}, the angular momentum in a disk is transported by the \emph{turbulent viscosity}. It seems to be appropriate to keep various possibility of the physical origin of viscosities. Hence, we treat the viscosities $\Pi$ and $\widetilde{\Pi}\,^{\mu\nu}$ as \emph{phenomenological} variables, whose physical origin is not specified. This means that the bulk and shear viscous rates, $\zeta$ and $\eta$, are left as undetermined parameter in the equations~\eqref{eq:heat} $\sim$~\eqref{eq:shear}. As mentioned in previous paragraph, $\zeta$ and $\eta$ are the phenomenological parameter whose values should be determined empirically. The others of EIT's basic equations are given by the conservation of mass current, $(\rho\,u^{\mu})_{;\mu}=0$, and that of stress-energy-momentum tensor, $T^{\mu\nu}_{\quad;\nu}=0$ ; \begin{eqnarray} \cd{\rho} + \rho\,u^{\mu}_{\,\,\,;\mu} &=& 0 \\ \rho\,\left( \cd{\varepsilon} + p\,\cd{V} \right) &=& - q^{\mu}_{\,\,\,; \mu} - q^{\mu}\,\cd{u}_{\mu} - \left(\, \Pi\,\Delta^{\mu\nu} + \widetilde{\Pi}\,^{\mu\nu} \,\right)\,u_{\mu\,;\,\nu} \\ \label{eq:eom} \left(\, \rho\,\varepsilon + p + \Pi \,\right)\,\cd{u}\,^{\mu} &=& - \cd{q}\,^{\mu} + q_{\alpha}\,\cd{u}\,^{\alpha}\,u^{\mu} - u^{\alpha}_{\,\,\,; \alpha}\,q^{\mu} - q^{\alpha}\,u^{\mu}_{\,\,\,; \alpha} - \Delta^{\mu\alpha}\, \left(\, (p+\Pi)_{, \alpha} + \widetilde{\Pi}\,^{\,\,\,\beta}_{\alpha\,\,\,; \beta} \,\right) \,, \end{eqnarray} where $\varepsilon$, $p$ and $T$ are the state variables of fiducial equilibrium state. The above equations~\eqref{eq:heat}~$\sim$ \eqref{eq:eom} and the Einstein equation are the basic equations of EIT. Finally let us make a comment of which the original works of EIT~\cite{ref:eit,ref:israel,ref:is,ref:hl_causality,ref:hl_applicable} are not aware: The EIT can not be applied to radiation fluids as shown in~\cite{ref:rad}. Non-equilibrium radiation fluids need a special treatment different from the other matters. However, although the inconsistency of EIT with radiation fluid has been revealed, no satisfactory non-equilibrium phenomenology of radiation fluid exits at present. \section{Dissipative Accretion Flow onto a Schwarzschild Black Hole} As a preliminary report, we show two theorems, without precise proof, on dissipative flows around a Schwarzschild black hole. Both theorems can be proven by solving the basic equations of EIT without solving the Einstein equation and fixing the metric on Schwarzschild spacetime. The use of fixed background metric means that our analysis does not include the self-gravity of the dissipative fluid. Therefore the theorems shown below are applicable to non-self-gravitating dissipative flows, i.e. to the dissipative fluid whose total energy is sufficiently less than the mass of central black hole, however ``non-self-gravitating'' dissipative fluid is usually assumed in Newtonian accretion disk theories~\cite{ref:disk}. For the first, let us consider the spherically symmetric dissipative accretion flow onto a Schwarzschild black hole: \begin{thm} Bulk and shear viscosity do not vanish in stationary spherically symmetric dissipative accretion flows onto Schwarzschild black hole. \end{thm} Naively one may think that the shear viscosity should vanish in spherical accretion flow. However theorem~1 denies this naive sense. Here note that the Navier-Stokes law also allows the existence of shear viscosity in spherical flow~\cite{ref:ray}. The difference between EIT and Navier-Stokes is that, while a flow wit only bulk (or shear) viscosity is possible in Navier-Stokes law, but such flow in EIT has an unphysical divergence; the density or speed of accretion diverges in the region far from black hole. Hence, in EIT, the bulk and shear viscosities do not arise separately, and they interact always. This result about viscosity may be a significant property of EIT in contrast with Navier-Stokes law. Now I am trying to extend this theorem to a non-stationary spherical dissipative flow on general non-stationary spherical spacetime which includes the self-gravity of dissipative fluid. For the second, let us consider the stationary toroidal dissipative flow around a Schwarzschild black hole: \begin{thm} For stationary toroidal flow with no mass accretion onto black hole, any flow of finite temperature is impossible, but only a rigid toroidal rotation of zero-temperature is possible. The fluid velocity and dissipative fluxes of the possible zero-temperature flow are \begin{equation} \label{eq:u} u^t = \frac{1}{\sqrt{f(r)-\Omega^2\,r^2\,\sin^2\theta}} \quad,\quad u^r = u^\theta = 0 \quad,\quad u^\phi = \Omega\,u^t \qquad\mbox{and}\quad q^{\mu}=\Pi=\widetilde{\Pi}\,^{\mu\nu}=0 \,, \end{equation} where $(t,r,\theta,\phi)$ is the Schwarzschild coordinate, $f(r) := 1 - 2M/r$ with $M$ as the black hole mass, and $\Omega$ is the angular velocity measured by a rest observer. Since $\Omega$ is constant and dissipations vanish even when they are switched on, the velocity $u^{\mu}$ in equation~\eqref{eq:u} expresses a rigid rotation. And the thermodynamic state variables of this flow are \begin{equation} \label{eq:state} T = p = \varepsilon = 0 \quad,\quad \rho(r,\theta) = \mbox{arbitrary in the region $f(r)>\Omega^2\,r^2\,\sin^2\theta$} \end{equation} Here recall that the internal energy $\varepsilon(x)$, for a fluid element at $x$, includes the mass energy, kinetic energy and the potential energy by external gravity due to black hole. Therefore the vanishing internal energy $\varepsilon=0$ means that, for each fluid element, the potential energy due to black hole cancels out the kinetic and mass energies of fluid element. \end{thm} This theorem together with the third law of thermodynamics implies that any flow of finite temperature is non-stationary and/or non-toroidal, and never relaxes to the flow given in equations~\eqref{eq:u} and~\eqref{eq:state}. Hence, although a differentially rotating and stationary toroidal flow of finite temperature is possible for ``perfect'' fluid~\cite{ref:fm}, such stationary flow becomes unstable once the dissipations are switched on. Theorem~2 implies the dissipative instability of an exact solution of relativistic perfect fluid. The dissipative instability, the EIT's effect, may describe a formation of out-flow from an accretion disk without introducing a magnetic field (so-called MRI), while the ``acceleration'' of out-flow to form a jet will be explained with magnetic fields. In (near) future, it is expected to apply the EIT to a plasma composing an accretion disk.
\section{Introduction} Ultra-faint satellites of the Milky Way include dwarf galaxies ~\citep{Wi05,Zu06a,Zu06b,Be06a,Be07,Ir07,Be08} and star clusters ~\citep{Ko07}, as well as objects with intermediate properties ~\citep{Wal07,Be09, Ni09}. Defined by their extremely low surface brightness, these systems could only be detected with a massive multi-band imaging campaign like the Sloan Digital Sky Survey (SDSS). In this {\it Letter}, we announce the discovery of two further Milky Way satellites in the Southern Galactic portion of the SDSS SEGUE survey. They lie in adjacent constellations and each extend only a couple of arc-minutes on the sky. However, their heliocentric distances differ by an order of magnitude, and, hence, so do their physical sizes. We name the dwarf galaxy in the constellation of Pisces, lying at the heliocentric distance of $\sim 180$ kpc and measuring $\sim 120$ pc across, Pisces II. This is the second Galactic stellar halo sub-structure in Pisces - the first, Pisces I was announced earlier this year by ~\citet{Wat09}. Pisces I is much closer and more dispersed on the sky: it is at least several degrees across and lies at $\sim 80$ kpc. Our second discovery is a feeble cluster of stars in the constellation of Pegasus. It has a half-light radius of 3 pc and lies at a heliocentric distance of 16 kpc. We name it Segue 3, after SEGUE, the imaging survey in the data of which it was found. In the analysis presented in this Letter we have extinction-corrected all magnitudes using the maps of \citet{Sc98}. \begin{figure*}[t] \begin{center} \includegraphics[width=0.9\textwidth]{segue3_pisces2_figure1_sdss.ps} \caption{The SDSS view of Pisces II (upper row) and Segue 3 (lower row): {\it Left:} Density of stars in the SDSS catalogue centered on the object. The stars in the $10^\prime \times 10^\prime$ area are binned into $15\times15$ bins and smoothed with a Gaussian with FWHM of 1.5 pixel. {\it Middle Left:} CMD of all stars in a circle of radius $1\farcm2$ (marked on the left panel), dominated by the satellite's members. {\it Middle Right:} Comparison CMD of stars within the annulus $7^\prime$ to $7\farcm1$ showing the foreground. {\it Right:} Difference in Hess Diagrams. Pisces II populations (top right) can be hinted at by the red giant branch and blue horizontal branch. Segue 3 (bottom right) shows an obvious main sequence. Ridge-lines of M92 (red, [Fe/H]$\sim$-2.3) and M13 (blue, [Fe/H]$\sim$-1.55) are over-plotted. For Segue 3, we also show a mask built using M92 ridge-line to select possible red giant stars for luminosity calculation (lower middle left panel).} \label{fig:fig1_sdss} \end{center} \end{figure*} \begin{deluxetable}{lcc} \tablecaption{Properties of Pisces II and Segue 3 \label{tbl:pars}} \tablewidth{0pt} \tablehead{ \colhead{Parameter} & {Pisces II} & {Segue 3}} \startdata RA (J2000) & $22:58:31 \pm 6$ & $21:21:31 \pm 4$ \\ Dec (J2000) & $+05:57:09 \pm 4$ & $+19:07:02 \pm 4$ \\ Galactic $\ell$ & $79.21^\circ$ & $69.4^\circ$ \\ Galactic $b$ & $-47.11^\circ$ & $-21.27^\circ$ \\ $r_h$ (Plummer) & $1\farcm1 \pm 0\farcm1$ & $0\farcm65 \pm 0\farcm1$ \\ $\theta$ & $77^\circ \pm 12^\circ$ & $215^\circ \pm 20^\circ$\\ $e$ & $0.4 \pm 0.1$ & $0.3 \pm 0.2$\\ (m$-$M)$_0$ & $21\fm3$ & $16\fm1$ \\ M$_{\rm tot,V}$ & $-5\fm0$ & $-1\fm2$ \enddata \tablenotetext{*}{Magnitudes are accurate to $\sim \pm 0\fm5$ and are corrected for the Galactic foreground reddening.} \label{tab:struct} \end{deluxetable} \section{Data and Discovery} The SDSS imaging data are available through the latest Data Release 7 (DR7) in two parts: i) $\sim 8000$ square degrees of the main SDSS field of view, mostly around the North Galactic Cap and ii) $\sim 3000$ square degrees of SEGUE imaging at low Galactic latitudes, with large portions of the Southern Galactic sky covered~\citep{Ya09}. These imaging data are produced in five photometric bands, namely $u$, $g$, $r$, $i$, and $z$ and are automatically processed through the same pipelines to measure photometric and astrometric properties \citep{Ab09}. The two imaging datasets differ in the continuity of the sky coverage and the amount of Galactic reddening: while most of the SDSS has contiguous coverage and is observed through minimal amounts of dust, SEGUE consists of tens of long, $2.5^\circ$-wide stripes affected by various amounts of Galactic extinction. Both discontinuity in coverage and variable extinction complicate the search for stellar over-densities by adding non-Poissonian noise to the stellar density field. Nonetheless, applying our over-density detection algorithm \citep{Be06a} to the SEGUE data immediately yielded several promising candidates, which we are continuing to follow up with deep imaging and spectroscopy. We have already presented the first result, the discovery of the Segue 2 satellite \citep{Be09}. Several more candidate objects were detected at similar significance level. Pisces II was selected for follow-up, as it showed tentative evidence for the presence of Blue Horizontal Branch (BHB) stars. The case for Segue 3 was simpler: the object could actually be seen on the SDSS images. Fig.~\ref{fig:fig1_sdss} shows the view of Pisces II and Segue 3 as seen by SDSS. The first of the four panels shows the density of all objects classified as stars by the SDSS pipeline down to $r=23$. In each case, there is a visible over-density at the center: Pisces II is detected with significance \citep{Ko08} of $\sim 5$ and Segue 3 with significance of $\sim 7$. The next two panels are the color-magnitude diagrams (CMDs) of the object and the Galactic foreground around it. It requires a lot of imagination to see the red giant branch or indeed the blue horizontal branch of Pisces II. However, the Hess difference in the right panel of Fig.~\ref{fig:fig1_sdss} shows that there are over-densities of blue-ish and reddish stars that can be interpreted as BHBs and RGBs at the heliocentric distance of $\sim 180$ kpc. The lower panel of Fig.~\ref{fig:fig1_sdss} presents the SDSS data for Segue 3, which is clearly a simpler case: the main sequence (MS) at $\sim 15$ kpc is obvious. It is, however, more difficult to identify the RGB population in Segue 3 as there is no obvious over-density in the Hess difference plot. In the absence of spectroscopic data, we gauge the likely membership by placing a mask that selects six potential members, at least one of which probably belong to the Galactic foreground. There is also a lone blue star in Segue 3, which looks slightly too faint to be classified as a BHB unambiguously. \begin{figure*} \begin{center} \includegraphics[width=0.9\textwidth]{segue3_pisces2_figure1_kpno.ps} \caption{The KPNO view of Pisces II (upper row) and Segue 3 (lower row). The panels are the same as in Figure 1, with the inner radius increased to $2^\prime$. The KPNO saturation limit is fainter than that of SDSS at $r\sim 18$, but it reaches $\sim 2$ magnitudes deeper. The CMD of Pisces II shows the most obvious improvement. In the KPNO data, strong RGB and clear BHB and blue straggler populations can now be seen. The dotted box in the middle left is used to pick out the likely BHB members. The dotted lines in the right panel outline the regions used to select the satellite members.} \label{fig:fig1_kpno} \end{center} \end{figure*} \begin{figure} \begin{center} \includegraphics[width=0.49\textwidth]{segue3_pisces2_density.ps} \caption{Density contours (black) of candidate member stars selected from the KPNO data centered on each satellite. {\it Left:} Pisces II is mapped out with RGB stars. {\it Right:} Segue 3 is mapped out with MS stars. Contour levels are 3, 5, 8, 10, 15 $\sigma$ (also 20 and 30 $\sigma$ for Segue 3) above the background. Red ellipses show Plummer isodensity contours corresponding to one and two half-radii. Blue dots mark the locations of the BHB candidate stars.} \label{fig:density} \end{center} \end{figure} \begin{figure} \begin{center} \includegraphics[width=0.49\textwidth]{segue3_image.ps} \caption{Color images covering $4^\prime \times 4^\prime$ region centered on Segue 3 made with SDSS (left) and KPNO (right) data. SDSS image is made with $g,r $ and $i$ band frames. KPNO image is made with $g$ and $r$ band frames.} \label{fig:seg_image} \end{center} \end{figure} \section{Follow-Up Imaging and Structural Parameters} During the nights of 19-24 September 2009, we obtained follow-up photometry of Pisces II and Segue 3 using the MOSAIC camera at the 4-m Mayall Telescope at Kitt Peak National Observatory, Arizona. For both objects, we observed a single $36\arcmin \times 36\arcmin$ field for $3\times 900$ seconds in both $g$ and $r$ filters. Each of MOSAIC's eight detectors have $2048 \times 4096$ pixels with scale $0.258\arcsec$ per pixel (unbinned). For this run we enjoyed photometric conditions and typical seeing of $\la 1\arcsec$. We processed raw MOSAIC frames using the set of IRAF-based procedures that have been developed for the NOAO Deep Wide-Field Survey\footnote{Details of these reduction procedures are available at http://www.noao.edu/noao/noaodeep/ReductionOpt/frames.html} \citep{Ja00}. These procedures include steps to correct for a pupil ghost that is caused by reflections off the atmospheric dispersion corrector. Briefly, we used averaged dome flats to generate a template image of the pupil ghost, which we then removed from the master flat field image. After dividing science frames by the master flat field, we then scaled and subtracted the pupil template image from science frames one by one. Visual inspection confirms that these steps successfully removed the pupil ghost from our science frames. Data stacking and the production of catalogues was performed using a general purpose pipeline for processing wide-field optical CCD data~\citep{Ir01}. For each image frame, an object catalog was generated and used to update the world coordinate system prior to stacking each set of 3 frames. A final set of object catalogs was generated from the stacked images and objects were morphologically classified as stellar or non-stellar (or noise-like). The detected objects in each passband were then merged by positional coincidence (within $1^{\prime\prime}$) to form a combined $g,r$ catalogue and photometrically calibrated on the SDSS system using stars in common. {\it Pisces II} -- The KPNO photometry reaches at least 2 magnitudes fainter than the SDSS and plays a crucial role in the identification and analysis of Pisces II, as can be seen from Fig.~\ref{fig:fig1_kpno}. The stellar over-density is enhanced, with a significance of $\sim 8$. The CMD is now quite unambiguous, with both RGB and BHB clearly visible together with a pile-up of stars around the main sequence turn-off (MSTO). The very tight BHB can now be used to estimate the distance modulus of the system (m$-$M)$_0 = 21.3$. Both CMD and the Hess difference panels of Fig.~\ref{fig:fig1_kpno} show over-plotted the ridge-lines of the globular clusters M92 and M13 from \citet{Cl05}. The RGB stars in Pisces II seem to be described equally well by both and, hence, are likely to possess metallicity in the range of $-2.3 < $[Fe/H]$ < -1.55$ \citep{Ha96}. Note, however, that the two brightest RGB stars are located right on top of the M13 ridgeline. To select all likely members of Pisces II, we draw a wide mask shown in the right panel of Fig.~\ref{fig:fig1_kpno}; we also select the BHB candidate stars with a color-magnitude box shown in the middle left panel. Using these selection cuts we can now measure the structural parameters of Pisces II and its luminosity. The density contours of the RGB and MSTO stars are shown in the left panel of Fig.~\ref{fig:density}. To measure the satellite's half-light radius, ellipticity and position angle, we follow the procedure outlined in \citet{Ma08}. From the locations of candidate MSTO and RGB stars we measure a half-light radius of $1\farcm1$ (or $\sim$ 60 pc at a distance of $\sim$ 180 kpc) with noticeable ellipticity. The results of this maximum likelihood fit are reported in Table~\ref{tab:struct}. To estimate the total luminosity of Pisces II, we i) integrate the flux inside the mask shown in the right-most panel of Figure~\ref{fig:fig1_kpno} within $3^{\prime}$ to get V$=16.6$ mag; ii) subtract an estimate of background contamination of $17.4$ mag; iii) add the contribution from the BHBs and possible blue stragglers (BS) of V $=19.4$ to get $M_{\rm V}=-4.2$. To account for the missing flux in the fainter stars, we use the luminosity function of the Ursa Minor dSph, which we integrate within $3.7 < $V$ < 13$ to get 0.8 mag. This can be compared to 0.5 mag correction if the luminosity function of a typical globular cluster is used (see e.g. \citet{Ni10}). Out of all ingredients contributing to the total luminosity, the exact number of RGB members of Pisces II carries the largest uncertainty. Our final estimate of the luminosity of Pisces II is $M_{\rm V}=-5\pm 0.5$ mag. {\it Segue 3} -- This satellite is one of the very few recent discoveries that can be seen seen directly in the SDSS images. Figure~\ref{fig:seg_image} shows color images of 4$^{\prime}\times4^{\prime}$ area centered on Segue 3 made with $g,r$ and $i$ SDSS frames and $g$ and $r$ KPNO frames. SDSS and KPNO data are of comparable quality, but the KPNO frames are integrated longer and hence the faint objects are detected and measured with greater accuracy. On both images, a central concentration of bluish stars belonging to Segue 3 can be seen, albeit with a somewhat irregular distribution. The KPNO photometry presented in the lower panel of Figure~\ref{fig:fig1_kpno} reveals a rather tight main sequence, with a clear turn-off, which we use to estimate the distance modulus of (m$-$M)$_0 = 16.1$. Of the two ridgelines over-plotted on the satellite's CMD, M92, with [Fe/H]$\sim$-2.3, clearly provides a better match. We use the KPNO data to derive the structural parameters of Segue 3 by following the procedure mentioned above. The half-light radius of Segue 3 is $0 \farcm 65$ and, overall, its distribution of stars is circular (see Table~\ref{tab:struct}), although there are some irregularities in the density profiles as can be seen from the right panel of Figure~\ref{fig:density} most likely due to the small numbers of stars. We calculate the luminosity of Segue 3 in two steps, using both SDSS and KPNO data. For the MS members fainter than $r=19$, we integrate the flux within $2^{\prime}$ inside the mask shown in the bottom right panel of Figure~\ref{fig:fig1_kpno} and subtract the estimate of the background contamination (outside $2^{\prime}$) to get V$=16.6$. For the RGB members ($r<19$), we integrate flux inside a narrower mask shown in the middle-left panel of Figure~\ref{fig:fig1_sdss}, which, after the background subtraction, gives V$=15.15$. So, the total V$=14.9$, or $M_{\rm V}=-1.2$. This is accurate to not less than 0.5 mag and can be better constrained when the true RGB members are identified through spectroscopic follow-up. \section{Discussion and Conclusions} Pisces~II is close on the sky to Pisces~I, discovered as an overdensity of RR Lyraes by \citet{Wat09} and confirmed spectroscopically by \citet{Ko09}. But, at a heliocentric distance of $\sim 180$ kpc, Pisces~II is almost twice as far as away. The extent of Pisces~I is probably considerable, at least as judged from the RR Lyrae populations (see Figure 12 of Watkins et al. 2009). This might imply the break-up of a substantial satellite galaxy moving on a radial orbit, in which case it is natural to interpret Pisces~II as a further fragment or companion. Pisces~II is very similar in morphology, size and luminosity to a number of recent discoveries such as Hercules, Leo~IV and Leo~V~\citep{Be07, Be08}. They also all lie at similar heliocentric distances of $\sim 150$ kpc. All four have extended BHB populations enshrouding them. Segue~3 is a very close relative of Koposov~1 and 2, the ultrafaint star clusters at distances of $\sim 50$ kpc~\citep{Ko07}. All three objects have a similar size ($\sim 3$ pc), luminosity $M_{\rm V}\sim -1$ and contain only a few tens of stars. Unlike Koposov~1 and 2, Segue~3 is much closer, and might even be a part of the Hercules-Aquila Cloud~\citep{Be07b,Wat09}. The evolution of such objects is known to proceed with pronounced mass segregation. The very few heavy stars sink to the centre, and the lighter stars are ejected to form a diffuse corona. It would be interesting to verify this prediction with spectrocopic surveys of the objects. \acknowledgments Funding for the SDSS and SDSS-II has been provided by the Alfred P. Sloan Foundation, the Participating Institutions, the National Science Foundation, the U.S. Department of Energy, the National Aeronautics and Space Administration, the Japanese Monbukagakusho, the Max Planck Society, and the Higher Education Funding Council for England. The SDSS Web Site is http://www.sdss.org/. VB thanks the Royal Society for financial support. SK, MAW, LLW and LW all thank the Science and Technology Facilities Council of the UK for funding. EO acknowledges NSF grant AST-0807498; MM acknowledges NSF grant AST-0808043
\section{Introduction} Strong bars are observed in optical images of roughly half of all the nearby disk galaxies \citep{Marinova2007, Barazza2008, Aguerri2009}. Therefore bars are a common feature in the central regions of disk galaxies. Their growth is partly regulated by the exchange of angular momentum with the stellar disk and the dark matter halo. For this reason the dynamical evolution of bars can be used to constrain the content and distribution of dark matter in the inner regions of galaxy disks \citep[e.g.,][]{Weinberg1985, Bertin1989, Debattista2000}. The morphology and dynamics of a barred galaxy depend on the angular velocity or pattern speed of the bar, $\Omega_{\rm bar}$. Usually, the bar pattern speed is parametrized with the bar rotation rate ${\cal R}\equivD_L/a_B$. This is the distance-independent ratio between the corotation radius $D_L$, where the gravitational and centrifugal forces cancel out in the rest frame of the bar, and the length of bar semi-major axis $a_B$. The corotation radius is derived from the bar pattern speed as $D_L = V_{\rm c}/\Omega_{\rm bar}$, where $V_{\rm c}$ is the disk circular velocity. As far as the value of ${\cal R}$ is concerned, if ${\cal R} < 1.0$ the stellar orbits are elongated perpendicular to the bar and the bar dissolves. For this reason, self-consistent bars cannot exist in this regime. Bars with ${\cal R} \gtrsim 1.0$ are close to rotating as fast they can, and there is no a priori reason for ${\cal R}$ to be significantly larger than 1.0. The value of ${\cal R}$ can be used to classify bars into fast ($1.0 \leq {\cal R} \leq 1.4$) versus slow (${\cal R} > 1.4$), with the dividing value at 1.4 by consensus \citep{Debattista2000}. Note the value of ${\cal R}$ does not imply a specify value of the pattern speed. \section{Indirect methods for measuring the bar pattern speed} A variety of indirect methods has been used to measure the pattern speed of bars and their corresponding rotation rates. The identification of rings with the location of Lindblad's resonances \citep[e.g.,][]{VegaBeltran1998} and analysis of the offset and shape of dust lanes which trace the location shocks in the gas flows \citep{Athanassoula1992, Puerari1997} promise to be simple and physically motivated methods to derive $\Omega_{\rm bar}$. But, they are based on the correct interpretation of morphological features which are often elusive. The comparison of the observed gaseous kinematics to dynamical models of gas flow \citep[e.g.,][]{Lindblad1996} and the comparison of the observed morphology to the predictions of numerical simulations \citep[e.g.,][]{Rautiainen2008} combine both kinematic and photometric information. Furthermore, dynamical models can be applied to highly-inclined systems. But, both methods usually do allow to put strong constraints on the error budget and solution uniqueness. In spite of being model-dependent, all the above methods give consistent results. The rotation rate of nearly all the measured bars is consistent to be $1.0 \leq {\cal R} \leq 1.4$ within the errors \citep{Elmegreen1996, Rautiainen2008}. However, the sample is biased toward late-type barred galaxies, because gas-rich systems are required for this kind of analysis. \section{The Tremaine-Weinberg method} \citet[][hereafter TW]{Tremaine1984} suggested a model-independent way for measuring the bar pattern speed. They showed that $\Omega_{\rm bar}$ can be determined from readily observable quantities for a tracer population satisfying the continuity equation. It is \begin{eqnarray} \lefteqn{\Omega_{\rm bar}\, \sin i =} \nonumber \\ & & \frac{\int^{+\infty}_{-\infty}\,h(Y)\,dY \int^{+\infty}_{-\infty}\,V(X,Y)\,\Sigma(X,Y)\,dX} {\int^{+\infty}_{-\infty}\,h(Y)\,dY \int^{+\infty}_{-\infty}\,X\,\Sigma(X,Y)\,dX}, \label{eq:tw} \end{eqnarray} where $(X,Y)$ are the Cartesian coordinates in the sky plane, with the origin at the disk center and the $X$ and $Y$ axes aligned with the disk major and minor axes, $h(Y)$ is an arbitrary weight function, $\Sigma$ and $V$ are the surface brightness and line-of-sight velocity of the tracer, and $i$ is the disk inclination. The integration in $X$ ranges over $-\infty \leq X \leq +\infty$, but integrating over $-X_0 \leq X \leq X_0 $ is sufficient if the disk is axisymmetric at $|X|\geq X_0$. Although the integration in $Y$ ranges over $-\infty \leq Y \leq +\infty$, it is actually performed over an arbitrary range because of $h(Y)$. For example, a weight function proportional to a delta function $\delta(Y-Y_0)$ corresponds to an aperture parallel to the disk major axis and offset by a distance $Y_0$. This is the case of the slits and pseudo-slits in long-slit and integral-field spectroscopy, respectively. \citet{Merrifield1995} refined the TW method. They suitably normalized both the numerator and denominator of the right-hand side of Eq. \ref{eq:tw} with the total luminosity in the aperture. Thus, the TW equation takes the form \begin{equation} \Omega_{\rm bar}\,\sin i = \frac{{\cal V}}{{\cal X}}, \label{eq:mk} \end{equation} where \begin{equation} {\cal X} = \frac{\int^{+\infty}_{-\infty}\,h(Y)\,dY \int^{+\infty}_{-\infty}\,X\,\Sigma(X,Y)\,dX} {\int^{+\infty}_{-\infty}\,h(Y)\,dY \int^{+\infty}_{-\infty}\,\Sigma(X,Y)\,dX} \label{eq:pin} \end{equation} and \begin{equation} {\cal V} = \frac{\int^{+\infty}_{-\infty}h(Y)\,dY \int^{+\infty}_{-\infty}V(X,Y)\,\Sigma(X,Y)\,dX} {\int^{+\infty}_{-\infty}\,h(Y)\,dY \int^{+\infty}_{-\infty}\,\Sigma(X,Y)\,dX} \label{eq:kin} \end{equation} are the luminosity-weighted means of the position and line-of-sight velocity of the tracer, respectively. Plotting ${\cal V}$ versus ${\cal X}$ for the different apertures produces a straight line with slope $\Omega_{\rm bar} \sin i$, where the inclination of the galaxy disk is known from the analysis of the surface-brightness distribution of the galaxy. The assumption underlying the TW method is that the observed surface brightness is proportional to the surface density of the tracer, as for old stellar populations in the absence of significant star formation and patchy obscuration of dust. Thus, the method has been successfully applied to absorption-line spectra of early- and intermediate-type barred galaxies (Table~1). Extensions of the TW method were proposed to determine the distinct pattern speeds of two nested bars within a single galaxy \citep{Corsini2003, Maciejewski2006} and to measure a pattern speed which may vary arbitrarily with radius \citep[][see also Meidt, this volume]{Merrifield2006}. Usually the presence of shocks, conversion of gas between different phases, and star formation on short timescales prevent the application of the TW method to gas. These limitations were explored by \citet{Rand2004} and \citet{Hernandez2005} with numerical experiments. Ionized \citep{Hernandez2004, Hernandez2005, Emsellem2006, Fathi2007, Fathi2009, Chemin2009, Gabbasov2009}, atomic \citep{Bureau1999}, and molecular gas \citep{Zimmer2004, Rand2004} were used to derive the pattern speed of the bar (and/or spiral arms) in a growing number of intermediate- and late-type barred galaxies. The agreement between the bar parameters derived from the gas-based TW method and those obtained from indirect methods and numerical simulations suggests that although the gas does not obey the continuity equation, it can be used to derive the bar pattern speed. Nevertheless, a detailed estimate of the systematic effects due to departures from continuity and a comprehensive comparison with stellar-based TW measurements (and corresponding bar rotation rates) are still missing. Moreover, TW measurements of the same galaxy based on different tracers have been not performed yet. In the rest of this paper, I will focus on the application of the TW method to absorption-line spectra, while John E. Beckman will review the results obtained from emission-line spectra in his contribution (this volume). \begin{table*} \begin{center} \caption{Barred galaxies with $\Omega_{\rm bar}$ measured by applying TW method to the stellar component} \begin{small} \begin{tabular}{llrrrrrl} \noalign{\smallskip} \hline \noalign{\smallskip} \multicolumn{1}{c}{Galaxy} & \multicolumn{1}{c}{Morp. Type} & \multicolumn{1}{c}{$D$} & \multicolumn{1}{c}{$a_{\rm bar}$} & \multicolumn{1}{c}{$\Omega_{\rm bar}$} & \multicolumn{1}{c}{$R_{\rm CR}$} & \multicolumn{1}{c}{$\cal{R}$} & \multicolumn{1}{c}{Ref.} \\ \multicolumn{1}{c}{} & \multicolumn{1}{c}{} & \multicolumn{1}{c}{(Mpc)} & \multicolumn{1}{c}{(kpc)} & \multicolumn{1}{c}{(km s$^{-1}$ kpc$^{-1}$)} & \multicolumn{1}{c}{(kpc)} & \multicolumn{1}{c}{} & \multicolumn{1}{c}{} \\ \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)} \\ \noalign{\smallskip} \hline \noalign{\smallskip} ESO 139-G09 & (R)SAB0$^0$(rs) & 71.9 & $5.9^{+2.2}_{-1.0}$ & $61\pm17$ & $5.1^{+1.8}_{-1.1}$ & $0.8^{+0.3}_{-0.2}$ & A$+$03\\ ESO 281-G31 & SB0$^0$(rs) & 70.1 & $3.7\pm0.3$ & $31\pm12$ & $6.8^{+4.1}_{-1.4}$ & $1.8^{+1.1}_{-0.4}$ & G$+$03\\ IC 874 & SB0$^0$(rs) & 34.7 & $3.3^{+0.9}_{-0.8}$ & $42\pm14$ & $4.5^{+2.2}_{-1.1}$ & $1.4^{+0.7}_{-0.4}$ & A$+$03\\ NGC 271 & (R$'$)SBab(rs) & 50.3 & $7.1\pm0.2$ & $32\pm18$ & $10.7^{+7.3}_{-3.9}$ & $1.5^{+1.0}_{-0.5}$ & G$+$03\\ NGC 936 & SB0$^+$(rs) & 14.9 & $3.6\pm0.4$ & $66\pm15$ & $5.0\pm1.1$ & $1.4^{+0.5}_{-0.4}$ & MK95\\ NGC 1023 & SB0$^-$(rs) & 5.8 & $1.9\pm0.1$ &$181\pm64$ & $1.5^{+1.0}_{-0.6}$ & $0.8^{+0.5}_{-0.3}$ & D$+$02\\ NGC 1308 & SB0/a(r) & 82.4 & $5.0^{+0.7}_{-1.4}$ & $99\pm35$ & $3.6^{+1.8}_{-0.9}$ & $0.8^{+0.4}_{-0.2}$ & A$+$03\\ NGC 1358 & SAB0/a(r) & 51.6 & $4.8\pm0.8$ & $37\pm18$ & $5.8^{+4.8}_{-1.8}$ & $1.2^{+1.0}_{-0.4}$ & G$+$03\\ NGC 1440 & (R$'$)SB0$^0$(rs):& 18.4 & $2.2\pm0.5$ & $82\pm19$ & $3.4^{+1.0}_{-0.6}$ & $1.6^{+0.5}_{-0.3}$ & A$+$03\\ NGC 2523 & SBbc(r) & 46.0 & $7.5\pm1.0$ & $30\pm7$ & $9.9^{+3.2}_{-1.9}$ & $1.3^{+0.7}_{-0.5}$ & T$+$07\\ NGC 2950 & (R)SB0$^0$(r) & 19.7 & $3.3\pm0.2$ &$117\pm25$ & $3.1^{+0.8}_{-0.6}$ & $1.0^{+0.3}_{-0.2}$ & C$+$03\\ NGC 3412 & SB0$^0$(s) & 16.0 & $2.4\pm0.2$ & $57\pm16$ & $3.6^{+1.3}_{-0.8}$ & $1.5^{+0.6}_{-0.3}$ & A$+$03\\ NGC 3992 & SBbc(rs) & 16.4 & $4.5\pm1.0$ & $72\pm5$ & $3.6\pm0.2$ & $0.8\pm0.2$ & G$+$03\\ NGC 4245 & SB0/a(r): & 15.6 & $2.9\pm0.4$ & $62\pm25$ & $3.2^{+2.2}_{-0.9}$ & $1.1^{+1.1}_{-0.4}$ & T$+$07\\ NGC 4431 & dSB0/a & 15.0 & $1.6\pm0.1$ &$102\pm26$ & $0.9^{+0.3}_{-0.2}$ & $0.6^{+0.2}_{-0.1}$ & C$+$07\\ NGC 4596 & SB0$^+$(r) & 29.3 & $7.5\pm1.1$ & $28\pm7$ & $8.6^{+2.8}_{-1.7}$ & $1.1^{+0.7}_{-0.3}$ & G$+$99\\ NGC 7079 & SB0$^0$(s) & 32.8 & $4.0\pm0.6$ & $53\pm1$ & $4.9\pm0.2$ & $1.2^{+0.3}_{-0.2}$ & DW04\\ \noalign{\smallskip} \hline \noalign{\smallskip} \noalign{\smallskip} \noalign{\smallskip} \end{tabular} \begin{minipage}{13.4cm} NOTE -- Col.(2): Morphological classification from \citet[][RC3]{RC3}, except for ESO 281-G31 (NASA/IPAC Extragalatic Database, NED) and NGC~4431 \citep{Barazza2002}. NGC 2950 is a double-barred galaxy and the listed values refer to its primary bar. Col.(3): Distance obtained as $V_{\rm CBR}/H_0$ with $V_{\rm CBR}$ from RC3, except for ESO 281-G31 (NED) and NGC~1308 (NED). The Hubble constant is assumed to be $H_0=75$ km~s$^{-1}$~Mpc$^{-1}$. Col.(4): Bar length from reference papers, except for NGC~936 \citep{Kent1989}. Uncertainties of $\pm14\%$ (corresponding to the median error of the remaining galaxies) are assigned to the bar lengths of NGC~2523, NGC~4245, and NGC~4596, since errors are not quoted in the reference papers. Col.(5): Bar pattern speed from reference papers. Col.(6): Corotation radius from reference papers. Errors for NGC~2523 and NGC~4245 are not given in \citet{Treuthardt2007}. Uncertainties are assigned according to the quoted errors on their pattern speeds. Col.(7): Bar rotation rate from reference papers. Uncertainties of NGC~936, NGC~2523, NGC~4245, and NGC~4596 are assigned according to quoted errors on bar length and corotation radius. Col.(8): Reference papers. \end{minipage} \end{small} \end{center} \end{table*} \section{Error budget} The main sources of uncertainties in TW measurements of $\Omega_{\rm bar}$ are summarized as follows. \smallskip \noindent {\it Centering errors:\/} The value of $\Omega_{\rm bar}$ can be significantly affected by small errors in identifying the position $(X_{\rm C},Y_{\rm C})$ of galaxy center and in measuring the value $V_{\rm sys}$ of systemic velocity. This effect was already recognized by \citet{Tremaine1984}. To counter it, they suggested to adopt a weight function which is odd in $Y$, since barred galaxies are nearly point-symmetric about their centers. In long-slit spectroscopy the centering problem translates into one of fixing an arbitrary reference position and velocity frame common to all the slits. This in general is a much easier task to achieve. To this aim \citet{Merrifield1995} rewrote Eq.~\ref{eq:mk} as \begin{equation} \Omega_{\rm bar}\,\sin i = \frac{{\cal V}-V_{\rm sys}}{{\cal X}-X_{\rm C}}. \end{equation} In integral-field spectroscopy the centering errors are minimized by the unambiguous determination of the common reference frame, which allows to know the exact position at which velocity and surface brightness of the tracer are measured. \smallskip \noindent {\it Signal-to-noise ratios:\/} ${\cal V}$ and ${\cal X}$ measure differences of velocity and luminosity across $X=0$, respectively and are suscep\-ti\-ble to the noise of available data. The signal-to-noise ratio of the spectral data can be increased by collapsing a long-slit spectrum along its spatial direction \citep{Merrifield1995} or by coadding all the spectra within a pseudo-slit \citep{Debattista2004}. This produces a single one-dimensional spectrum with a high signal-to-noise ratio. The mean Doppler-shift of its absorption lines gives the ${\cal V}$ value, which is required by the TW method. Broad-band luminosity profiles have higher signal-to-noise ratios than luminosity profiles derived from spectra, particularly at large radii. Therefore, they can be adopted to compute the value of ${\cal X}$ at each position of the slits or pseudo-slits \citep{Aguerri2003}. \smallskip \noindent {\it Uncertainties on the disk position angle:\/} The TW method requires that the slits (or pseudo-slits) be exactly parallel to the disk major axis. A careful determination of the disk position angle PA$_{\rm disk}$ is therefore required before placing the slits or extracting the pseudo-slits. The maximum permitted misalignment between the position angle of the slits (or pseudo-slits) and PA$_{\rm disk}$ to have reliable $\Omega_{\rm bar}$ measurements depends on the galaxy inclination and bar orientation with respect the line of nodes \citep{Debattista2003}. It ranges from $1^\circ$ to $4^\circ$ for $\Delta \Omega_{\rm bar}/\Omega_{\rm bar}=0.3$. Galaxies with an inclination of about $60^\circ$ and a bar at about $20^\circ$ from the line of nodes on the disk plane are less sensitive to misalignment. The analysis of the galaxy isophotes mapped by deep imaging of the axisymmetric region of a disk gives PA$_{\rm disk}$ and $i$. Imaging is also useful to identify and discard target galaxies with structures (e.g., outer rings, spiral arms, warped or non-axisymmetric disks) which may interfere with the accurate measurement of PA$_{\rm disk}$ and $i$ and affect the determination of $\Omega_{\rm bar}$. \smallskip \noindent {\it Dust obscuration and star formation:\/} The TW method requires that the observed surface brightness of the tracer be proportional to its surface density. This is almost strictly satisfied by old stars in early-type disk galaxies, because they are characterized by slow star formation rate and low content of dust. \citet{Gerssen2007} investigated the effects of dust obscuration and star formation on the stellar-based TW measurements by means of numerical simulations. They find that $\Delta \Omega_{\rm bar}/\Omega_{\rm bar} = 0.05$ for a diffuse disk of dust with a typically observed value of extinction $A_V = 3$. A unrealistically large $A_V=8$ gives $\Delta \Omega_{\rm bar}/\Omega_{\rm bar} \leq 0.15$. In addition, barred galaxies often display prominent dust lanes which run along the leading edges of the bar from the end of the bar toward the center of the galaxy \citep[see][for a morphological classification of dust lanes]{Athanassoula1992}. Dust lanes tend to increase the TW-derived value of $\Omega_{\rm bar}$ when the position angle of the bar with respect to the disk major axis is PA$_{\rm bar} > 0^\circ$ and decrease it when PA$_{\rm bar} < 0^\circ$. It is $0.08 \leq \Delta \Omega_{\rm bar}/\Omega_{\rm bar} \leq 0.25$ for realistic dust lanes with $A_V\simeq3$. Including star formation does not affect these conclusions. These experiments show that it is possible to extend the application of the TW method to the stellar component of late-type barred galaxies (see also Gerssen \& Debattista, this volume). The effects of dust obscuration could be further minimized by performing near-infrared spectroscopy. \smallskip \noindent {\it Number of slits (or pseudo-slits):\/} The value of $\Omega_{\rm bar} \sin{i}$ is derived by a straight-line fit to ${\cal X}$ and ${\cal V}$ data. They are measured along different slits (or pseudo-slits) crossing the galaxy bar and parallel to the disk major axis. Therefore, the accuracy of $\Omega_{\rm bar}$ determination depends also on the number of the observed slits (or pseudo-slits). It ranges from 2 \citep[NGC~1358, $\Delta \Omega_{\rm bar}/\Omega_{\rm bar}=0.49$;][]{Gerssen2003} to 9 (slits for the primary bar of NGC~2950, $\Delta \Omega_{\rm bar}/\Omega_{\rm bar}=0.21$; \citealt{Corsini2003}; pseudo-slits for NGC~7079, $\Delta \Omega_{\rm bar}/\Omega_{\rm bar}=0.02$; \citealt{Debattista2004}). The most common case corresponds to 3 slits, they were measured for half of the sample galaxies. The comparison between $\Delta \Omega_{\rm bar}/\Omega_{\rm bar}$ of NGC~2950 and NGC~7079 show that $\Delta \Omega_{\rm bar}$ is dominated by the uncertainties on ${\cal X}$ and ${\cal V}$. The median relative error on $\Omega_{\rm bar}$ is $\Delta \Omega_{\rm bar}/\Omega_{\rm bar}=0.27$. \smallskip TW measurement of $\Omega_{\rm bar}$ requires no modeling. However, in the absence of gas velocities at large radii, the determination of ${\cal R}$ requires some modeling to recover the disk circular velocity. The sources of uncertainties on ${\cal R}$ are discussed hereafter. \smallskip \noindent {\it Uncertainties on the bar length:\/} Determining the length of a bar is not a entirely trivial task. This is particularly true for SB0 galaxies, for which there is no spiral structure or star formation beyond the bar marking its end. Moreover, the presence of a large bulge complicates further the measurement of $a_B$ \citep{Aguerri2005}. Several methods have been developed to derive the bar length \citep[see][for a list]{Aguerri2009}. Methods used to measure $a_B$ for the galaxies in Table 1 include: Fourier decomposition of galaxy light to analyze bar-interbar intensity ratio \citep{Debattista2002, Aguerri2003, Gerssen2003, Corsini2003, Corsini2007, Debattista2004} or phase angle of Fourier mode $m=2$ \citep{Aguerri2003, Gerssen2003, Corsini2003, Corsini2007}, study of the radial profile of ellipticity \citep{Debattista2002} or phase angle of the deprojected ellipses which best fit galaxy isophotes \citep{Debattista2002, Aguerri2003, Corsini2003, Corsini2007}, identification of a change in the slope of the surface-brightness profile along the bar major axis \citep{Gerssen1999}, visual inspection of galaxy images \citep{Treuthardt2007}, and decomposition of the surface-brightness distribution \citep{Kent1989, Kent1990, Aguerri2003, Gerssen2003, Corsini2003, Corsini2007}. The relative error on the bar length for galaxies with at least two independent measurements is $\Delta a_B/a_B<0.25$ with a median $\Delta a_B/a_B=0.14$. For each galaxy $\Delta a_B$ is assumed to be the average of the error intervals at $68\%$ confidence level. \smallskip \noindent {\it Uncertainties on the corotation radius:\/} The corotation radius is obtained from the bar pattern speed and disk circular velocity. In the absence of gas which traces the circular velocity at large radii, $V_{\rm c}$ is recovered from the observed streaming velocities, velocity dispersions, and light distribution of the stellar component by applying the asymmetric drift correction \citep[see][p. 354]{GD2ed}. The difference between the stellar and circular velocity can be fairly large in the disks of bright SB0 galaxies ($\Delta V/V=0.2$), where large stellar velocity dispersions ($\sigma\simeq100$ km~s$^{-1}$) are observed \citep{Debattista2002, Aguerri2003, Corsini2003}. Relative error $\Delta V_{\rm c}/ V_{\rm c}$ ranges between 0.05 and 0.2 (with $\Delta V_{\rm c}$ defined as the average of $68\%$ error intervals). It includes the scatter of the observed velocities in the flat portion of the stellar rotation curve and variation of the parameters adopted for the asymmetric drift correction. Uncertainties on the bar pattern speed and disk circular velocity translate into a relative error on the corotation radius as large as $\Delta D_L/D_L=0.57$ \citep[NGC~1358;][]{Gerssen2003}. The median relative error of the sample is $\Delta D_L/D_L=0.28$ (with $\Delta D_L$ average of $68\%$ error intervals). \begin{figure*}[] \begin{center} \resizebox{10cm}{!}{\includegraphics[clip=true]{corsini_f01.ps}} \end{center} \caption{\footnotesize The corotation radius $D_L$ as a function of the bar length $a_B$ for the barred galaxies with $\Omega_{\rm bar}$ measured by applying the stellar-based TW method. Red circles denote measurements for SB0 galaxies, blue diamonds correspond to SB0/a galaxies, and green squares represent spiral barred galaxies. The dotted lines correspond to ${\cal R}=1$ and ${\cal R}=1.4$, respectively. They separate the forbidden (${\cal R}<1$), fast-bar ($1.0\leq{\cal R}\leq1.4$, hatched area), and slow-bar (${\cal R}>1.4$) regimes. The dashed line marks ${\cal R}=2$.} \label{fig:r} \end{figure*} \section{Discussion and conclusions} All the bars in Table~1 are consistent with being fast (Fig.~\ref{fig:r}). This is particularly true when galaxies with small uncertainties on ${\cal R}$ are considered. In fact, the probability that the bar length is twice as long as the corotation radius (${\cal R}>2$) is $12\%$ for all the sample galaxies and $8\%$ for galaxies with $\Delta {\cal R} / {\cal R}\leq0.3$. The median rotation rate is ${\cal R}=1.2$. The fact that some of the values of bar rotation rate are nominally ${\cal R}<1$ has been interpreted by \citet{Debattista2003} as due to the scatter introduced by uncertainties on PA$_{\rm disk}$. The quoted uncertainties on ${\cal R}$ are heterogeneous and include both $68\%$ confidence intervals and maximal errors. It would be very useful if they were calculated and given in an homogeneous way. For example, they could be estimated from Monte Carlo simulations based on the uncertainties on $a_B$ and $D_L$. No trend in ${\cal R}$ is observed with morphological type. But, the sample of bars studied so far with the stellar-based TW method is biased toward the bright and strongly barred SB0 and SB0/a galaxies. One of them hosts two nested bars (NGC~2950, \citealt{Corsini2003}). The sample includes only 3 spiral galaxies (NGC~271, \citealt{Gerssen2003}; NGC~2523, \citealt{Treuthardt2007}; NGC~3992, \citealt{Gerssen2003}), and one dwarf galaxy (NGC~4431, \citealt{Corsini2007}). The bar rotation rate of NGC~4431 is ${\cal R} = 0.6^{+1.2}_{-0.4}$ at $99\%$ confidence level \citep{Corsini2007}. Albeit with large uncertainty, the probability that the bar ends close to its corotation radius is about twice as likely as that the bar is much shorter than the corotation radius. This suggests a common formation mechanism of the bar in both bright and dwarf galaxies. If their disks were previously stabilized by massive dark matter halos, bars were not produced by tidal interactions because they would be slowly rotating \citep{Noguchi1999}. But, this is not the case even in the two sample galaxies which show signs of weak tidal interaction with a close companion (i.e., NGC~1023, \citealt{Debattista2002}; NGC~4431, \citealt{Corsini2007}). There is no difference between TW measurements of the stellar component in isolated or mildly interacting barred galaxies. Besides, neither the length nor strength of the bars are found to be correlated with the local density of the galaxy neighborhoods \citep{Aguerri2009}. The bars of ESO~139-G09 \citep{Aguerri2003} and NGC~1358 \citep{Gerssen2003} are weak and fast. Thus, the hypothesis of \citet{Kormendy1979} that weak bars are the end state of slowed down fast bars is not supported by observations. They instead favor a scenario in which weak and strong bars form in the same way. The ${\cal R}$ determinations based on the stellar TW method agree with those obtained by indirect methods, which are largely adopted for gas-rich galaxies. According to the compilations by \citet{Elmegreen1996} and \citet{Rautiainen2008}, almost all the measured bars have $1.0 \leq {\cal R} \leq 1.4$ within the errors. The same is true also for the ${\cal R}$ values obtained from pattern speeds measured with the gas-based TW method \citep{Bureau1999, Fathi2009, Chemin2009}. A fast bar is ruled out by errors only in NGC~2917 \citep[${\cal R}>1.7$,][]{Bureau1999} and UGC~628 \citep[${\cal R}=2.0^{+0.5}_{-0.3}$,][]{Chemin2009}. If bars of gas-poor lenticulars and early-type spirals have the same ${\cal R}$ as gas-rich late-type spirals, then gas is not dynamically important for the evolution of bar pattern speed \citep{Debattista2003}. Unfortunately, uncertainties on the measured ${\cal R}$ are often not quoted for late-type galaxies. This missing piece of information is crucial to derive the ${\cal R}$ distribution as a function of the morphological type. Additional work is needed for the stellar TW method to increase both the accuracy of the $\Omega_{\rm bar}$ measurements and extend the number of studied late-type barred galaxies. Integral-field spectroscopy overcomes many problems of long-slit observations and leads to more efficient and accurate TW measurements. Dwarf barred galaxies are ideal targets to this aim. They nicely fit the field of view of the available integral-field spectrographs and are good candidates for testing the predictions that dark-matter dominated barred galaxies should have a slow bar. A successful application of the TW method to the stellar component of late-type barred galaxies would remedy the selection bias present in the current sample of measured $\Omega_{\rm bar}$. This will allow a straightforward comparison of the results of stellar-based TW method with indirect and gas-based TW measurements in the same range of Hubble types. \begin{acknowledgements} I would like to thank Victor Debattista for careful reading of the manuscript and comments that helped to improve it. I would also like to thank Alfonso Aguerri, John Beckman, and Lorenzo Morelli for their suggestions. This research has been made possible by support from Padua University through grant CPDR095001. It has made use of the Lyon Extragalactic Database (LEDA) and NASA/IPAC Extragalactic Database (NED). \end{acknowledgements} \bibliographystyle{aa}
\section{Introduction} The environmental dependence of galaxy properties, such as broadband color, star formation, and stellar mass, is a well-known effect in the local universe \citep{Dressler-80}. In this context environment means an estimate of the smoothed density filed at a given location. Of particular interest for the present study is the dependence of the galaxy stellar mass function (GMF) on environment. Recent comprehensive galaxy redshift surveys have led to intensive studies in this field. For instance, \cite{Mo-04} model the dependence of the luminosity function on the large-scale environment based on mock catalogs for the the two-degree Field Galaxy Redshift Survey \citep[2dFGRS;][]{Colless-01} and \cite{Baldry-06} investigate the GMF as a function of environment in the nearby universe based on the Sloan Digital Sky Survey \citep[SDSS;][]{York-00}. To uncover evolutionary effects, surveys spanning a larger redshift range have been used: the study by \cite{Bundy-06} is based on the DEEP2 Galaxy Redshift Survey ($0.4\leq z \leq 1.4$); \cite{Pannella-09} use the COSMOS survey; \cite{Bolzonella-09} employ the zCOSMOS survey in the redshift range $0<z<1$; and \cite{Scodeggio-09} investigate the environment dependence of the GMF based on the VVDS survey covering a redshift range of $0.2 < z < 1.4$. On intermediate scales, $\sim\hMpc$, groups and clusters of galaxies themselves provide a definition of environment. At low redshift, \cite{Balogh-01} using Two Micron All Sky Survey (2MASS) and Las Campanas Redshift Survey (LCRS) data, separate different environments as field, groups, and clusters, finding that galaxy luminosity and mass functions depend on both, galaxy type (with steeper functions for emission line galaxies) and mass of the group (with more massive and brighter objects more common in clusters), mainly as a consequence of the different contribution of passive galaxies (see also \citealt{Hansen-09}). The environmental dependence of properties of gas is linked to the studies of warm--hot intergalactic media (WHIM). Hydrodynamical simulations by \cite{Cen-Ostriker-99} show that the average temperature of baryons is an increasing function of time, with most of the baryons at the present time having a temperature in the range of $10^5-10^7\rm K$. The detection of this warm-hot gas poses an observational challenge. While according to their census more than one-half of the normal matter is yet to be detected, it is not clear which methods have to be used. In a later study \cite{Cen-Ostriker-06} report that the gas density of the warm-hot intergalactic medium is broadly peaked at a density three to seven times the critical density, however their dark matter mass resolution was only moderate and a question of assigning baryons to groups has not been addressed. All the above studies would greatly benefit from a quantification of the contribution of galaxy groups to the density fields as a function of background density. The significance of this question has been recognized before \citep[e.g.,][]{Sheth-Tormen-99}, but a detailed answer has never been provided. Simulation work has instead concentrated on questions regarding halo assembly bias and, in particular, on changes in the properties of galaxies within the halos of similar mass but residing in different environments \citep{Lemson-Kauffmann-99, Gao-Springel-White-05, Croton-Gao-White-07}. This paper is organized as follows. In Section~\ref{sec:data} we review the Millennium Simulation, the semi-analytic modeling of galaxies and the determination of the background density. Section~\ref{sec:massfrac} examines the dependence of the dark matter halo mass fractions on environment. In Section~\ref{sec:massfunc} halo mass functions in different environments are discussed. Section~\ref{s:stars} focuses on the dependence of the GMFs on environment. We present a short summary in Section~\ref{sec:conclusion}. \begin{figure*} \epsfig{file=mfield.063.256.g10.041.eps,width=0.495\hsize} \epsfig{file=FoF_NPmfield.063.256.g10.041.eps,width=0.495\hsize}\\ \epsfig{file=mfield.041.256.g10.041.eps,width=0.495\hsize} \epsfig{file=mfield.027.256.g10.041.eps,width = 0.495\hsize} \caption{\label{fig:step} Upper left panel: dark matter density field at $z=0$, smoothed using a Gaussian filter with a smoothing scale of $10\hMpc$, $\delta_{\rm G10}$, and scaled by the average rms fluctuations within the corresponding volume, $\sigma_{\rm G10}$. Upper right panel: density field as shown on the left in gray scale. The points represent halos above $10^{12}{\>h^{-1}\rm M}_\odot$ within a slice of $10\hMpc$ thickness. Colors (from green to red) and sizes (from small to large) correspond to the logarithmic mass of the halos. As expected, the halo distribution closely follows the density field. Lower panels: smoothed and scaled density fields,as the upper left panel, but for the redshifts $z=1$ and $z=3$. According to linear theory $\delta/\sigma$ does not change with time.} \end{figure*} \begin{figure} \epsfig{file=ldd_mfieldg10.eps,width = 0.95\hsize} \caption{\label{fig:dd} Differential mass distribution as a function of scaled density contrast. The density is computed based on a $10\hMpc$ Gaussian smoothing kernel, and the scaling factor, $\sigma_{\rm G10}$, is the mass variance within a corresponding volume. Solid, dotted, and dashed lines indicate the redshifts $z=0$, $1$ and $3$, respectively. The red curves show the lognormal fits to the distribution.} \end{figure} \begin{figure} \epsfig{file=bg_fof.eps, width = 0.95\hsize} \caption{\label{fig:bg} Upper panel: mean CiC density as a function of smoothed density. Lower panel: mean density within isodensity contours divided by that isodensity.} \end{figure} \section{Data} \label{sec:data} This analysis is based on the publicly available Millennium simulation run \citep[MS; ][]{Springel-05a, Lemson-Springel-06}. In the first part of this section, we review the MS and the semianalytic galaxy modeling. The second part describes the determination of background densities and halo mass fractions. \subsection{Millennium simulation} \label{sec:mille} The MS adopts concordance values for the parameters of a flat $\Lambda$ cold dark matter cosmological model, $\Omega_{\rm dm}= 0.205$ and $\Omega_{\rm b}= 0.045$ for the current densities in cold dark matter and baryons, $h= 0.73$ for the present dimensionless value of the Hubble constant, $\sigma_8= 0.9$ for the rms linear mass fluctuation in a sphere of radius $8 \hMpc$ extrapolated to $z= 0$, and $n= 1$ for the slope of the primordial fluctuation spectrum. The simulation follows $2160^3$ dark matter particles from $z=127$ to the present day within a cubic region $500\hMpc$ on a side. The resulting individual particle mass is $8.6\times10^8{\>h^{-1}\rm M}_\odot$. The gravitational force has a Plummer-equivalent comoving softening of $5{\>h^{-1}\rm kpc}$. The Tree-PM $N$-body code GADGET2 \citep{Springel-05b} has been used to carry out the simulation and the full data are stored 64 times spaced approximately equally in the logarithm of the expansion factor. The halos are found by a two-step procedure. In the first step, all collapsed halos with at least 20 particles are identified using a standard friends-of-friends (FoF) group-finder with linking parameter $b = 0.2$. These objects will be referred to as FoF-halos. Then, post-processing with the substructure algorithm SUBFIND \citep {Springel-01} subdivides each FoF halo into a set of self-bound {\it sub-halos}. Based on their assembly histories, individual sub-halos are populated with galaxies by a semi-analytic prescription and various observable quantities are generated. For a detailed description of the semi-analytic galaxy catalog we refer the reader to \cite{Croton-06} and \cite{DeLucia-Blaizot-07}. The stellar mass functions in this study are based on the DeLucia2006a\_SDSS2MASS catalog \citep[][{\tt http://www.g-vo.org/Millennium}] {Lemson-VirgoConsortium-06}. The following analysis is based on FoF halos which hereafter are addressed simply as halos. In principle, one also could use different halo definitions like those derived from spherical top-hat overdensity criteria or gravitational self-boundedness. The former would ease a comparison with observations and the latter would not suffer from the bridging problem inherent to the FoF approach. However, according to various tests which we have performed, the results based on the various halo identification schemes show little difference. The advantage of using FoF halos lies in the simplicity of the approach which makes it a very common tool for the analysis of $N$-body simulations. \subsection{Background Density and Halo/Galaxy Mass Fraction} \label{sec:density} The MS database also provides information on the global density field besides describing individual halo properties. Here, we use the densities, $\rho_{\rm CiC}$, which are based on a Counts in Cell (CiC) approach using cubic cells of $~\sim2\hMpc$ on a side \citep[][]{Hockney-Eastwood-88} and $\rho_{\rm G10}$ (hereafter {\it background density}), which are derived from the former by smoothing them with a $10\hMpc$ Gaussian kernel. In this study, we utilize $\rho_{\rm CIC}$ to compute the total amount of matter in a given volume and calibrate the use of the $10\hMpc$ background density fields, $\rho_{\rm G10}$. We assume that $\rho_{\rm G10}$ represents the linear density field. To determine the total mass of halos per cell, we re-implement the CiC approach only taking into account the particles belonging to FoF halos above a given mass limit. If a halo crosses the boundary of a cell, only the mass of the halo inside the cell is counted. This quantity is used to compute the halo {\it mass fractions} in Section~\ref{sec:haback}. For the determination of the halo and the GMFs (Section~\ref{sec:massfunc} and Section~\ref{s:stars}), we attribute the total mass of halos or galaxies to the cell within which their centers are located. Eventually, each cell is assigned a total mass in halos, a total mass in stars (galaxies), a total mass, and a smoothed background density (which after multiplication with the cell volume corresponds to a mass as well). The upper left and the two lower panels of Figure~\ref{fig:step} display a slice of the density contrast, $\delta_{\rm G10} = (\rho_{\rm G10}-\langle\rho\rangle) / \langle\rho\rangle$ scaled by the variance of the matter fluctuations within a volume corresponding to the Gaussian filter, $\sigma_{\rm G10}$, at redshifts $z=0$, $z=1$, and $z=3$. Typical patterns of the large-scale density field are apparent, such as roughly spherical high density regions, filaments, and voids. According to linear theory, $\delta/\sigma$ does not change with time. Indeed, the scaled density contrast is very similar in shape and amplitude at the redshifts we consider. Figure~\ref{fig:dd} shows the differential mass distributions as functions of the scaled density contrast. Red curves show the lognormal fits to the distribution \citep[cf.,][]{Coles-Jones-91, Neyrinck-Szapudi-Szalay-09}. The approximate invariability of the scaled contrast reaffirms the presumption of linearity of $\delta_{\rm G10}$. The top right panel of Figure~\ref{fig:step} shows the halo distribution within a slice of $10\hMpc$ thickness at $z=0$ at the same location as slice used for the contrasts. The under-laid gray scale image is a replication of the contrast on the left. As expected, the halo distribution follows the pattern of the background density field. In the following, we measure the dependence of the fraction of mass captured in halos or galaxies as a function of the background density. \begin{figure*} \epsfig{file=m3_fofg10.eps, width = 0.95\hsize} \caption{\label{fig:maiha} Halo to total mass fraction as a function of background density contrast , $\delta_{\rm G10}$. As indicated, the line styles correspond to different lower mass limits for the halos used to compute to total halo mass budget. For the left and middle panels, these masses correspond to the equivalent peak heights $\nu= 0.9$, $1.0$, $1.2$, $1.4$, $1.7$, and $2.1$. In the $z=3$ panel on the right only the limiting masses corresponding to $\nu = 1.4$, $1.7$, and $2.1$ are shown. At that redshift, $\nu$-values below $1.4$ correspond to halos with less than $\sim20$ particles which are not resolved.} \end{figure*} \begin{figure} \epsfig{file=iso_fof.eps, width = 0.95\hsize} \caption{\label{fig:iso512} Conditional cumulative halos mass function, $F(\nu)$, i.e., fraction of mass locked in halos with respect to the total mass within a region confined by the contrasts, $\delta_{\rm G10}=0.5$, $1.0$, and $1.5$ as a function of halo mass given parameterized by the equivalent peak height $\nu = \delta_{\rm c}(z)/\sigma(M,z)$. The mass scale at the top indicates the equivalent masses for $z=0$. Solid, dotted, and dashed lines correspond to redshifts $z=0$, $1$, and $3$, respectively. At $z=3$ $\delta_{\rm G10}$ does not reach $1$.} \end{figure} \section{Halo mass fractions} \label{sec:massfrac} The first part of this section notes some general features associated with the characterization of the density field. In the subsequent paragraph, we examine the dependence of the halo mass fraction on environment. The {\it halo mass fraction} is defined as the mass locked in halos divided by the total mass in a given volume. As stated before, the environment is quantified based on the smoothed density within this volume. The halo mass fraction is closely related to the cumulative halo mass function which will be discussed in the subsequent section. \subsection{Mass within Isodensity Surfaces} \label{sec:iso} The upper panel in Figure~\ref{fig:bg} shows the ratio between the average CiC-density of all cells located in regions of a given density contrast, $\delta_{\rm G10}$, as a function of the density contrast itself. Line styles correspond to different redshifts. At a contrast of $\delta_{\rm G10}\approx0.2$, the mass within a cell approximately corresponds to the value of $\rho_{\rm G10}$ multiplied by the cell volume. For higher contrasts, the actual mass deposited in the cell is larger than that deduced from the smoothed density field. The opposite is true for contrasts below $0.2$. This behavior simply reflects that extremes are leveled by the smoothing procedure. More important for the subsequent analysis is the lower panel in Figure~\ref{fig:bg}. It displays the mean value of $\rho_{\rm CIC}/\rho_{\rm G10}$ within a volume confined by the isodensity surface, $\rho_{\rm G10}$, as a function of $\rho_{\rm G10}$ or it's equivalent $\delta_{\rm G10}$. The ratio, $\rho_{\rm CIC}/\rho_{\rm G10}$, gives the average factor with which the confining density has to be multiplied to get the true mean density or the true mass within the enclosed volume. For different redshifts, indicated by different line styles, one finds slightly different ratios, but the overall behavior is similar. For density contrasts between $0$ and $1.5$, which are of interest here, one obtains $1\lesssim\rho_{\rm CIC}/\rho_{\rm G10}\lesssim3$. This factor has to be accounted for when the total mass within a given volume is inferred from the confining isodensity surface. Loosely speaking, a factor of 2 has to be multiplied to the value of the confining surface density, $\rho_{\rm G10}$, to recover the mass inside. \subsection{The dependence of the halo mass fraction on environment} \label{sec:haback} The three panels in Figure~\ref{fig:maiha} show the {\it halo mass fractions} for the redshifts $z=0$, $z=1$, and $z=3$. The halo mass fraction is defined as the fraction of mass locked in halos (above a given limiting mass) and the total mass within a volume confined by isodensity surfaces with a density contrast $\delta_{\rm G10}$ as a function of that contrast. Line styles refer to halo mass limits as indicated. For each redshift, these mass limits correspond to fixed values of the equivalent peak height, $\nu(M,z) = \delta_{\rm c}(z)/\sigma(M,z)$, where $\sigma(M,z)$ is the rms linear overdensity within a sphere containing the mass $M$ at redshift $z$, and $\delta_{\rm c}(z)$ is the linear overdensity threshold for collapse at that redshift. The usage of $\nu$ instead of the actual halo mass should remove much of the cosmology dependence of our results and will be most useful for the discussion of the halo mass functions below. The mass limits for the left and middle panels correspond to $\nu\approx 0.9$, $1.0$, $1.2$, $1.4$, $1.7$, and $2.1$. In the $z=3$ panel (on the right) mass resolution only allows us to show the graphs for the three highest $\nu$ values, $1.4$, $1.7$, and $2.1$. Figure~\ref{fig:iso512} shows the fraction of mass in halos with respect to the total mass for confining density contrasts of $\delta_{\rm G10} = 0.5$, $1.0$ and $1.5$ as a function of the mass limit of the halos. It is formally identical to the conditional cumulative mass function, $F(\nu)$, which is discussed in the next section. However, here it is computed based on the particle distributions of halos. Therefore, the mass of halos which transgress cell boundaries is accurately split between the cells. The analysis in the next section is somewhat more coarse in the sense that it attributes the total mass of a halo to only one cell, namely the cell where the center of the halo is located. This can introduce some bias, similar in origin to the deviation of surface and mean enclosed density discussed in ~\ref{sec:iso}. However, the comparison between the two methods yields acceptable agreement which implies confidence in the computation of the mass functions presented below. \subsection{Where are the Warm--Hot Baryons?} Our findings have consequences for the detection of warm-hot intergalactic baryons (WHIM). Cosmological simulations predict that some 50\% of all the baryons locally appear in the form of gas with temperatures between $10^5$ and $10^7\rm K$ \citep{Cen-Ostriker-99}. The main process responsible for heating the baryons to such temperatures is large scale structure formation and thus the missing component is thought to be associated with high densities. Observational detection of this component became subject of a number of studies, and yet a full account of warm--hot baryons has not been reached. In a subsequent study \cite{Dave-01} have argued that the major fraction of WHIM is located outside galaxy groups. Using hydrodynamical simulations they have shown that the total fraction of WHIM associated with galaxy groups is between 10\% and 25\%, depending on the group mass cut and that the majority of the WHIM resides at mean densities of $10<\delta<200$. In this work, no attempt has been undertaken to study the contribution of baryons in galaxy groups to the WHIM as a function of environment. Here we aim to determine the fraction of WHIM located in groups at high background densities. Figure~\ref{fig:maiha} suggests a dominant contribution of groups to the matter budget at high overdensities, resolving 60\% of total mass in groups with mass exceeding $2\times10^{13}M_\odot$ and 70\% of total mass in groups with mass exceeding $4\times10^{12}M_\odot$. Without any background density restriction these groups account for 20\%--30\% of matter. It is important to note that no gas component has been included in the Millennium Simulation. To derive conclusions on the WHIM, we rely on a model describing the distribution of gas relative to the overall matter distribution which is equivalent to dark matter distribution in the current context. Here, we assume for simplicity that gas follows dark matter. Hydrodynamical simulations show that this assumption is well justified for local densities below $\lesssim10^4$ times the cosmic mean density \citep[cf.,[]{Faltenbacher-07}. Furthermore, observational and theoretical accounts for baryons inside groups indicate that the universal baryon fraction is nearly recovered \citep{Kravtsov-Nagai-Vikhlinin-05,Sun-09,Giodini-09,McGaugh-10}, and potentially can be fully resolved by accounting the baryons near the virial radius. Thus, on the supposition that gas follows the dark matter our findings indicate that 20\%--30\% of the WHIM is located in groups if no background density constraint is imposed, whereas 60\%--70\% of the WHIM is residing in galaxy groups at high background densities. The former agrees well with the figures quoted in \cite{Dave-01}. The only observational result published on the global fraction of mass in groups, by \cite{Reiprich-Bohringer-02}, uses {\it ROSAT} All Sky Survey data, and quoted $\Omega_{cluster}=0.012$ (or 5\% contribution of clusters to the total mass budget) for masses exceeding $10^{14} M_\odot h^{-1}$. This measurement compares well with our value for $M>2\times 10^{14} {\>h^{-1}\rm M}_\odot$, which corresponds to the same $\nu$ when a difference in the $\sigma_8$ value between observed universe and the Millennium run is taken into account. This account will likely improve soon since current X-ray surveys can access masses below $10^{13}M_\odot$ \citep{Finoguenov-07, Finoguenov-09}. Also, detections of WHIM in X-ray emission regions agree well with our findings at high background densities. For example, \cite{Werner-08} find that most masses of baryons in the A222/A223 complex are locked within the halos, with WHIM contributing 10\%--20\%. At this point we would like to emphasize that the definition of densities adopted in WHIM studies is different from that one used here. Our results constrain the WHIM aspect within the definition of density fields typical to spectroscopic surveys \citep[e.g.,][]{Kovac-10}. In contrast the computation of the local density of the Lyman$_\alpha$ absorbers can only be derived from their HI column densities, resulting in a non-trivial role played by the absorber's size and shape. In \cite{Penton-Stocke-Shull-04} the first comparison to LSS density has been provided, showing the half of the absorbers reside in voids. Figure~\ref{fig:maiha} limits the importance of the WHIM component in regions of high density (as defined on a 10 Mpc scale) which is not associated with groups to 30\%. Therefore, the situation is much more favorable toward detecting missing baryons in underdense regions, but there the temperature of the gas will be quite low. In fact the major success in resolving the missing baryons has been achieved using techniques looking for colder gas \citep{Penton-Stocke-Shull-04}. \section{Cumulative halo mass functions} \label{sec:massfunc} Up to this point, we have used the particle distribution of the halos to split their mass among the cells they occupy. In the following we use the halo position to attribute the entire halo mass to one cell. This reduces the accuracy of the determination of the mass fraction slightly ($\lesssim 5\%$). However, that way the treatment of halos and model galaxies which are examined hereafter and have unknown spatial extent can be matched. Plenty of studies have been devoted to investigate the mass function of halos in the cosmological context. Early analytical approaches suggested that the mass function is universal, i.e.\ its shape is independent of time and background cosmology if adequate variables are used. In a seminal study, \cite{Press-Schechter-74} used the equivalent peak height, $\nu = \delta_{\rm c}/\sigma$ \citep[see also,][]{Bond-91, Lee-Shandarin-98, Sheth-Tormen-99}. Such models can help gain insight into the statistics of gravitational collapse. However, these highly nonlinear processes are complex, and a final validation of the models can only be obtained by direct comparison to numerical simulations. Based on $N$-body simulations, \cite{Jenkins-01} and \cite{Evrard-02} presented fitting formulae for the differential mass functions accurate to $\sim10\%-20\%$. These studies supported the view that mass functions are indeed universal. Also, the results obtained by \cite{Lukic-07} are consistent with a universal mass function; however, they report a mild redshift dependence at low redshifts \citep[see also][]{Reed-03}. \cite{Warren-06} found a fitting formula accurate to $\sim5\%$ for a fixed cosmology at $z=0$. Recently, however, based on a large set of $N$-body simulations with different cosmological parameters \cite{Tinker-08} revealed that the mass function can not be represented by a universal mass function at this level of accuracy. In particular they found that the amplitude of the mass function decreases monotonically by $\approx 20\%-50\%$ from $z=0$ to $z=2.5$. Several studies explore the mass functions at high redshifts \citep[e.g.,][]{Reed-07, Cohn-White-08} but they do not particularly focus on universality. In general, analytical and numerical studies preferentially discuss the (unconditional) differential mass function. For our purposes the integrated form, i.e.\ the unconditional cumulative halos mass function (UHMF) is more useful since it directly gives the fraction of mass in halos relative to the total mass. Additionally, the functional forms of the unconditional differential mass functions, as given in the literature, are complex and their integrals are even more so. Thus, for our purposes it seems to be a better strategy to directly examine the cumulative mass function and derive a simple fitting function for it. Consequently, in the first part of this section we introduce a fitting function for the UHMFs. In the second part, this function is modified to be applicable for the cumulative mass functions at different background densities, hereafter referred to as conditional cumulative halo mass functions (CHMFs). \begin{figure} \epsfig{file=sfE_FoF.eps, width = 0.95\hsize} \caption{\label{fig:sf} Upper panel: cumulative mass functions for redshifts $z=0.00$, $0.14$, $0.36$, $0.69$, $1.17$, $1.91$, and $3.06$ shown as solid, dotted, dashed, dot-dashed, three-dot-dashed, and solid lines, respectively. Halo mass is given parametrically through the equivalent peak height $\nu=\delta_{\rm c}/\sigma$. At the top the corresponding mass scale for $z=0$ is indicated. The red line presents a fit for the mass function at $z=0$. The fitting function including the parameters is quoted in the bottom line. Lower panel: residuals of the mass functions and the fit at $z=0$. The red line displays the residual for the mass function at $z=0$ which has been used for fitting. An accuracy of $\lesssim5\%$ is achieved for halo masses between $10^{10}$ and $10^{15}{\>h^{-1}\rm M}_\odot$.} \end{figure} \begin{figure*} \epsfig{file=a6_FoF_NPg10E.eps, width = 0.95\hsize} \caption{\label{fig:alion} Dependence of the halo mass fraction on environment for the redshifts $z=0$, $1$, and $3$. Lines styles correspond the CHMFs in regions confined by the density contrasts, $\delta_{\rm G10}$ as indicated. Green lines represent a fits to the UHMF using Eq.~\ref{equ:fit}. The fitting parameters are displayed at the bottom of the upper panels. Red lines show fits for the CHMFs using the fitting procedure described in Section~\ref{sec:cmf}. The only difference between the upper and lower panels is the mass normalization. In the upper panels, $F$ gives the halo mass fraction with respect to the total mass in the box and in the lower panels with respect to the total mass enclosed in the overdense regions.} \end{figure*} \begin{figure} \epsfig{file=cotran_FoF_NP063g10E.eps, width = 0.95\hsize} \caption{\label{fig:cotran} Coordinate transformation used to convert unconditional into conditional mass functions for the indicated density contrasts. To guide the eyes the green line for $\delta_{\rm G10}=0$ displays the identity. For higher contrasts, this transformation keeps the mass function for large values of $\ln (\nu)$ unchanged yet it causes the mass function at low $\ln (\nu)$ to level off.} \end{figure} \begin{figure} \epsfig{file=allstar_063g10.eps, width = 0.95\hsize} \caption{\label{fig:allstar} Environmental dependence of the cumulative galaxy mass function. Except for the usage of galaxy masses instead of halo masses the plot is equal to the upper left panel in Figure~\ref{fig:alion}; the line styles are adopted from there as well.} \end{figure} \begin{figure*} \epsfig{file=a8_063g10.eps, width = 0.95\hsize} \caption{\label{fig:a8} Upper panel: cumulative galaxy mass function in galaxy groups in different environments. Lower panel: differential galaxy mass function in galaxy groups in different environments. On each panel, we denote both the density field and the number of embedded groups.} \end{figure*} \begin{figure*} \epsfig{file=a2_063g10.eps, width = 0.95\hsize} \caption{\label{fig:a2} Differential galaxy mass function in different environments (black lines) separating the contribution of central (red lines) and non-central galaxies (green lines) in groups above the masses indicated in the upper right corner of the panels. The green lines show the contribution of galaxies not belonging to those groups. The particular role of central galaxies is that they provide a hump-like contribution to the mass function and explain a variation of that hump with the environment.} \end{figure*} \subsection{Unconditional Cumulative Halo Mass Functions} \label{sec:umf} Under the assumption that the initial density field is a Gaussian random field and the validity of the spherical collapse model, \cite{Press-Schechter-74} derived the mass fraction locked in halos above a given mass to the total mass in the universe, \begin{equation} \label{equ:PS} F_{\rm PS}(\nu) = \mathrm{erfc}\left({\nu\over\sqrt{2}}\right), \end{equation} where $\nu$ is the equivalent peak height as discussed in \$~\ref{sec:haback}, which, for a given cosmology, can be uniquely transformed into a corresponding halo mass, $M$. $F(\nu)$ is equivalent to the UHMF as introduced above. \cite{Press-Schechter-74} argued that this parameterization makes the mass function universal, i.e. independent of evolutionary changes and cosmological parameters which are covered by the time and cosmology dependence of $\nu$. Over the last 20 years, high-resolution N-body simulations demonstrated that Equation~\ref{equ:PS} reproduces the numerical halo mass function qualitatively. However, certain systematic deviations became apparent \citep[e.g.,][]{Sheth-Tormen-02, Warren-06, Tinker-08}. Compared to $N$-body results, Equation~\ref{equ:PS} produces too many halos with masses corresponding to $\nu=1$ and to few at the high mass end. Nevertheless, we use the functional form of Equation~\ref{equ:PS} as template for our fitting formula: \begin{equation} \label{equ:fit} F_{\rm fit}(\nu) = A\ \mathrm{erfc}\left({a\ \nu^{\ b}}\right), \end{equation} with the fitting parameters $A$, $a$, and $b$. The upper panel of Figure~\ref{fig:sf} displays UHMFs for FoF halos derived from the MS. Different line styles represent UHMFs derived from snapshots at redshift $z=0.00$, $0.14$, $0.36$, $0.69$, $1.17$, $1.91$, and $3.06$. The red line shows a fit to the UHMF at $z=0$. The corresponding fitting parameters are given in the panel. For the computation of the halo masses, we applied the correction formula proposed by \cite{Warren-06}: $N{\rm corrected} = N(1-N^{0.6})$, where $N$ denotes the number of particles within a given FoF halo. This correction is most effective at the low mass end. The fitting range is confined by FoF halos with masses between $2\times10^{10}{\>h^{-1}\rm M}_\odot$ (200 particles) and $10^{15}{\>h^{-1}\rm M}_\odot$. The lower panel of Figure~\ref{fig:sf} shows the residuals between the UHMFs at the given redshifts and the fit based on the $z=0$ mass function. The red line highlights the residual between the fit and the $z=0$ mass function. At $z=0$ and for masses in the range between $10^{10}{\>h^{-1}\rm M}_\odot$ and $10^{15}{\>h^{-1}\rm M}_\odot$, the fit is accurate to the 5\% level. However, we note an increasing offset with redshift which results in a deviation of 15\% at $z=3.06$. Similar findings have been reported in \cite{Tinker-08}. The expression, Equation~\ref{fig:sf}, is an excellent fitting function for the cumulative mass function at $z=0$. It is ``universal'', i.e., independent of cosmology at a degree to which the parameterization by the equivalent peak height, $\nu$, permits. \subsection{Conditional Cumulative Halo Mass Function} \label{sec:cmf} The term ``conditional mass function'' has been used to address two related problems: (1) to describe the mass distribution of progenitor halos which end up in a given halo at later times; (2) to refer to the mass distribution of halos within a region of a given background density \citep[cf.,][]{Mo-White-96, Sheth-98, Sheth-Lemson-99b, Sheth-Tormen-02} . Here we will investigate the latter. According to excursion set theory, the basic equation to approach such problems is \begin{equation} \label{eq:excurs} F(M|\delta_0,\sigma_0) = \mathrm{erfc}\left({1\over\sqrt{2}} {\delta_{\rm c}-\delta_0\over\sigma-\sigma_0}\right), \end{equation} where $\delta_{\rm c}$ and $\sigma$ are the same as used in Equation~\ref{equ:PS} and $\delta_0$ and $\sigma_0$ are the linear density contrast and the dispersion of the background density field. It gives the fraction of mass in collapsed halos of mass greater than $M$ (corresponding to $\nu=\delta_{\rm c}/\sigma$) in a region that has a linear density contrast $\delta_0$ \citep[for a comprehensive review including references see,][]{Zentner-07}. In the current context, two difficulties arise when this equation is to be applied. First, the boundary surface and hence the volume of the overdense regions can be quite irregular which makes it difficult to determine $\sigma_0$ appropriately. Second, to facilitate the comparison with observations, we intend to compute the mass fraction within regions {\it above} a given $\delta_{\rm lim}$, in other words cumulative with respect to the background density. However Equation~\ref{eq:excurs} gives the halo mass fraction {\it at} a given $\delta_0$. Therefore, obtaining an expression suitable for our purposes would require an integration of Equation~\ref{eq:excurs} with respect to $\delta_0$. This integration leads to a lengthy expression, which is too complex to be used as a model for a fitting formula (as done for the UHMF in Section~\ref{sec:umf}). Therefore, we choose a more phenomenological approach starting with the inspection of Figure~\ref{fig:alion}. Black lines display UHMFs and CHMFs in regions above a given density contrast at redshifts of $z=0$, $z=1$, and $z=3$. The CHMFs are determined by summing up the mass of halos which reside in cells above a given contrast $\delta_{G10}$. The difference between the upper and lower panels is only in normalization. In the upper panels, we use the total mass in the box whereas in the lower panels we use the total mass within the cells above $\delta_{G10}$. Therefore, the CHMFs in the upper panels lie systematically below the UHMFs; this is because a fraction of the total volume is excluded by the density criterion -- and so are the halos in it -- but the mass fraction is still computed with respect to the total mass in the box. In the lower panels, the CHMFs lie systematically above the UHMFs which is a result of the decreasing total mass in the volume above the given background density. The latter is more physical but the former illustrates the fact that at the high-mass end all CHMFs display the same behavior as the UHMF. The green lines show fits to the UHMF. The values of the fitting parameters are given at the bottom of the upper panels (here the UHMFs are fitted separately for each redshift). The red lines are fits to the CHMFs. \begin{table} \begin{center} \begin{tabular}{|c|c|c|c|c|c|} \hline Redshift&$\delta_{\rm G10}$&$\Delta$&$a$&$b$&$f_M$\\\hline $z=0$& 0.00 & 1.00 & 1.00 & 1.00 & 1.00 \\ & 1.00 & 0.58 & 7.76 & 0.28 & 1.51 \\ & 1.25 & 0.36 & 7.90 & 0.41 & 2.18 \\ & 1.50 & 0.24 & 8.45 & 0.55 & 3.30 \\ & 1.75 & 0.16 & 8.91 & 0.66 & 5.12 \\ & 2.00 & 0.11 & 9.77 & 0.76 & 8.16 \\ & 2.25 & 0.08 & 10.25 & 0.85 & 13.17 \\ & 2.50 & 0.06 & 10.87 & 0.92 & 20.55 \\\hline $z=1$& 0.00 & 1.00 & 1.00 & 1.00 & 1.00 \\ & 1.00 & 0.61 & 7.58 & 0.51 & 1.63 \\ & 1.25 & 0.34 & 7.28 & 0.71 & 3.34 \\ & 1.50 & 0.20 & 7.79 & 0.90 & 8.41 \\ & 1.75 & 0.12 & 8.69 & 1.05 & 23.86 \\ & 2.00 & 0.07 & 8.72 & 1.17 & 67.96 \\ & 2.25 & 0.05 & 10.18 & 1.27 &180.57 \\\hline $z=3$& 0.00 & 1.00 & 1.00 & 1.00 & 1.00 \\ & 1.00 & 0.74 & 7.37 & 0.91 & 1.77 \\ & 1.25 & 0.32 & 7.27 & 1.20 & 13.43 \\ & 1.50 & 0.12 & 6.18 & 1.41 &266.83 \\\hline \end{tabular} \caption{\label{tab:cmf} CHMF fitting parameters, $\Delta$,$a$,$b$ and mass factor $f_{m}$ for the background densities, $\delta_{\rm G10}$, and redshifts $z=0$, $1$, and $3$.} \end{center} \end{table} Guided by the behavior seen in the upper panels of Figure~\ref{fig:alion}, we conceive a fitting formula for the CHMFs which is based on a parameter-dependent coordinate transformation, $T(\nu)$, such that for an adequate set of parameters $F(T(\nu))$ (here $F$ denotes the UHMF) matches a given CHMF. We use the following functional form for these transformations, constructed to map high $\nu$ values onto themselves but narrowing the range for low $\nu$s: \begin{equation} \label{equ:tran} \ln (T(\nu)) = \Delta + {a x + (1.0-a)\ \ln[1.0+\exp(x)]\over b}, \end{equation} where $x = b\ (\ln(\nu)-\Delta)$. The values for the three parameters $\Delta$, $a$, and $b$ are listed in Table~\ref{tab:cmf} with the corresponding fits shown as red lines in Figure~\ref{fig:alion}. Figure~\ref{fig:cotran} illustrates the coordinate transformations at $z=0$. $\Delta$ determines where the deviation from one-to-one correspondence takes place, $a$ gives the left hand side slope and $b$ defines the smoothness of the transition between the two slopes. So far, we have described how to fit the CHMFs in the upper panels of Figure~\ref{fig:alion}. To get the mass normalization right, i.e., to transform the fits of the upper panels into the those of the lower panels, these $F(T(\nu))$ have to be multiplied by the mass factor $f_M$ which is listed in the rightmost column of Table~\ref{tab:cmf}. It gives the ratio of the total mass in the box to the mass confined to the considered overdense regions. Our findings may help to illustrate some of the main mechanisms shaping the halo mass functions in different environments. The top heavy shape of the CHMFs is caused by a relative lack of small mass halos at high background densities. For our way of parameterizing the background density, namely, by using isodensity surfaces to indicate all the volume interior to it, the shape of the cumulative mass functions at the high mass end is independent of environment. The only difference is induced by the normalization which reflects the mass confined to the high density regions. It causes the amplitude of the mass function to rise. As a concluding remark we note that the use of cumulative instead of differential mass functions proved valuable to extract these results. First, because it directly gives the fraction of mass in halos above a given mass with respect to the total mass. In addition, it's shape is simple compared to the differential mass function which eases finding a suitable fitting strategy. Finally, it allows us to easily derive some of its basic properties by analytically. For instance, it is obvious that the ``wrongly normalized'' CHMFs in the upper panels must coincide at the high mass end. Similarly easy to derive is that the fits in the lower panels must be confined by $1$ even for extrapolations to smallest halo masses. The fact that this is obviously not the case for two of the $z=3$ fits demonstrates the limits of the fitting procedure. However, in general the fits behave well, i.e., the amplitudes remain $<1$ even if extrapolated to small masses. \section{Conditional galaxy mass function} \label{s:stars} The semi-analytical modeling of galaxies included in the MS database \citep{DeLucia-Blaizot-07,Lemson-VirgoConsortium-06} provides stellar masses for each model galaxy. This allows us to compute the conditional cumulative galaxy stellar mass function (CGMF) in exactly the same way as the CHMFs have been determined, namely by summing up the mass of all galaxies which reside in cells above a given contrast $\delta_{G10}$. We set the lower limit for the galaxy masses to $10^9{\>h^{-1}\rm M}_\odot$. Figure~\ref{fig:allstar} shows the results of the CGMF normalized by the total mass in the box, which is the reason why all mass functions coincide at the high-mass end, equivalently to the behavior of the CHMFs. In general, the changes in CGMF are qualitatively very similar to that of the CHMFs: becoming top heavy at high background densities. The qualitative similarities between the CHMFs and CGMFs suggest that the dependence of the model galaxy mass function on environment is caused by the corresponding dependence of the halo mass function on environment, rather than on direct impact of environment on galaxy evolution (generally referred to as nurture effects). To test this conjecture, Figure~\ref{fig:a8} shows the model galaxy mass functions at different background densities with the additional restriction that they reside in dark matter halos of a given mass range as indicated by the labels right on top. The upper panels show the cumulative galaxy mass functions and the lower panels depict the GMF in differential form. The line styles correspond to different density thresholds listed in the figure. The integer numbers indicate the number of halos in those regions. Evidently, these numbers show a strong dependence on environment assuring that the galaxy mass functions displayed in a single panel are based on very different sets of host halos. Nevertheless, the resulting mass functions deviate by less than $10$\% for galaxy masses $\lesssim 10^{11}{\>h^{-1}\rm M}_\odot$. Therefore, the large changes in the CGMFs can almost entirely be accounted for by changes in the host halo population. Nurture effects may have a minor impact on shaping the model galaxy mass function in different environments. There are some shortcomings in the semianalytical model of galaxy formation used here \citep{DeLucia-Blaizot-07}. The two most important are: (1) cooling is instantaneously shut down for galaxies whose halo enters a larger one; (2) tidal forces are not allowed to strip off stars reducing the luminosity of a given galaxy. Nevertheless, these processes do mostly affect satellite galaxies. Thus, we believe that the behavior seen at the high mass end of the CGMFs should be a reliable prediction from the semianalytical model. To provide a further illustration, we display the differential galaxy mass function as a function of environment separating the contribution of central and non-central galaxies in groups in Figure~\ref{fig:a2}. The particular role of central galaxies is that they provide a hump-like contribution to the mass function and illustrate the variation of that hump with the environment, therefore explain another important observational result \citep[e.g.,][]{Bolzonella-09}. Explaining the hump-like feature as being due to the contribution of central galaxies (and hence a bimodal GMF with centrals and satellites as its two constituents) is somewhat different than the explanation given by \citet{Bolzonella-09}, namely that the hump is due to the contributions of red galaxies to the total GMF. In our view, the hump is a consequence of halo assembly, not of the transformation of blue into red galaxies, although the two processes might be linked in the sense that the assembly of central galaxies might also lead to quenching of their star formation in some halos. We therefore predict that the hump will also be seen in the blue galaxy mass function alone, not only in the total GMF, since many halos host giant (blue) spiral galaxies as their central galaxy. Indeed, this bimodality in the blue galaxy mass function has been observed by \cite{Drory-09}. The thermodynamical state of the accreting gas is expected to vary with redshift, enabling a cold accretion at $z>2$ and subsequent star formation in the central galaxy even in halos as massive as $10^{13} M_\odot$ \citep{Dekel-09}. These are expected to be the highest peaks of density field at those redshifts and therefore obey our predictions for the behavior in high density environments. Observational search for the transitional halo mass between the cold and the hot accretion mode is difficult, as this mass is well below the sensitivity of both spectroscopic and X-ray surveys for defining the galaxy groups. Instead, here we propose to use the shape of the galaxy mass function to determine at which halo mass the transition occurs. Since the central galaxies of groups make a hump-like contribution to the galaxy mass function, the blue central galaxies should constitute a blue hump. The location of the blue hump in the galaxy mass function can therefore be used through a comparison to numerical simulations to determine the transition halo mass scale for shutting down the star formation or environmental dependence of cold accretion mode in galaxy formation. In contrast to clustering studies, the proposed method does not induce a requirement on the data to be representative, and can therefore be applied to a field of any size or even a selected object, like a high-redshift supercluster. \section{Conclusions} \label{sec:conclusion} Using the Millennium Simulation and current schemes for density field reconstruction, we parameterize the conditional halo mass function. As an application, we consider the role of halos in explaining the missing baryon problem and the environmental dependence of galaxy mass functions. We show that in high-density environments galaxy groups provide a major contribution to total matter content. We discuss the implication of this result for search of missing baryons. Under the well-justified assumption that baryons follow dark matter, we show that its amount can be constrained using the observations of galaxy groups. We also point out that the environmental changes in galaxy mass functions are caused by changes in mass function of groups and, in particular, that a hump-like features in galaxy mass function is produced by the central galaxies of groups. \section*{Acknowledgments} The authors are thankful to the anonymous referee for insightful suggestions, to Simon White for valuable comments on the paper and to Eyal Neistein for helpful discussions. Andreas Faltenbacher is supported by the SA SKA bursay program and acknowledges the kind hospitality at South African Astronomical Observatory. Alexis Finoguenov acknowledges support from {\it Spitzer} UDS Legacy program to UMBC. The Millennium Simulation databases used in this paper and the web application providing online access to them were constructed as part of the activities of the German Astrophysical Virtual Observatory. The authors thank Gerard Lemson for his help with MS database and the comments on the manuscript.
\section{#1} \setcounter{equation}{0}} \usepackage[mathscr]{euscript} \renewcommand{\theequation}{\thesection.\arabic{equation}} \newcommand{\operatornamewithlimits{\overline{lim}}}{\operatornamewithlimits{\overline{lim}}} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}[theorem]{Lemma} \newtheorem{proposition}[theorem]{Proposition} \newtheorem{corollary}[theorem]{Corollary} \theoremstyle{definition} \newtheorem{assumption}{Assumption}[section] \newtheorem{definition}{Definition}[section] \theoremstyle{remark} \newtheorem{remark}{Remark}[section] \newtheorem{example}{Example}[section] \newcommand\cbrk{\text{$]$\kern-.15em$]$}} \newcommand\opar{\text{\raise.2ex\hbox{${\scriptstyle | }$}\kern-.34em$($} } \newcommand{\text{\rm tr}\,}{\text{\rm tr}\,} \makeatletter \def\dashint{% \operatorname% {\,\,\text{\bf--}\kern-.98em\DOTSI\intop\ilimits@\!\!}} \makeatother \newcommand{\WO}{\overset{\scriptscriptstyle0}% {W}\,\!} \newcommand\gb{\mathfrak{b}} \newcommand\gw{\mathfrak{w}} \newcommand\bR{\mathbb{R}} \newcommand\bP{\mathbb{P}} \newcommand\bB{\mathbb{B}} \newcommand\bD{\mathbb{D}} \newcommand\bL{\mathbb{L}} \def\mathbb{M}{\mathbb{M}} \newcommand\bQ{\mathbb{Q}} \newcommand\bS{\mathbb{S}} \newcommand\bW{\mathbb{W}} \newcommand\cB{\mathcal{B}} \newcommand\cF{\mathcal{F}} \newcommand\cI{\mathcal{I}} \newcommand\cJ{\mathcal{J}} \newcommand\cP{\mathcal{P}} \newcommand\cR{\mathcal{R}} \newcommand\cQ{\mathcal{Q}} \newcommand\cD{\mathcal{D}} \newcommand\cZ{\mathcal{Z}} \newcommand\cH{\mathcal{H}} \newcommand\cL{\mathcal{L}} \newcommand\cM{\mathcal{M}} \newcommand\cK{\mathcal{K}} \newcommand\cN{\mathcal{N}} \newcommand\cS{\mathcal{S}} \newcommand\cW{\mathcal{W}} \newcommand\frD{\mathfrak{D}} \newcommand\ul{\underline{l}} \newcommand\bi{\mathbf{i}} \newcommand\bj{\mathbf{j}} \newcommand\sfB{{\textsc b}} \renewcommand{\boldsymbol}[1]{\mathbf{#1}} \begin{document} \title[Kalman-Bucy filter and SPDEs]{ Kalman-Bucy filter and SPDEs with growing lower-order coefficients in $W^{1}_{p}$ spaces without weights} \author[N.V. Krylov]{N.V. Krylov}% \thanks{The work was partially supported by NSF grant DMS-0653121} \address{127 Vincent Hall, University of Minnesota, Minneapolis, MN, 55455, USA} \email{<EMAIL>} \dedicatory{Dedicated to D.L. Burkholder} \subjclass[2000]{60H15, 93E11} \keywords{Stochastic partial differential equations, Kalman-Bucy filter, Sobolev spaces, growing coefficients} \begin{abstract} We consider divergence form uniformly parabolic SPDEs with VMO bounded leading coefficients, bounded coefficients in the stochastic part, and possibly growing lower-order coefficients in the deterministic part. We look for solutions which are summable to the $p$th power, $p\geq2$, with respect to the usual Lebesgue measure along with their first-order derivatives with respect to the spatial variable. Our methods allow us to include Zakai's equation for the Kalman-Bucy filter into the general filtering theory. \end{abstract} \maketitle \mysection{Introduction} We consider divergence form uniformly parabolic SPDEs with bound\-ed VMO leading coefficients, bounded coefficients in the stochastic part, and possibly growing lower-order coefficients in the deterministic part. We look for solutions which are summable to the $p$th power, $p\geq2$, with respect to the usual Lebesgue measure along with their first-order derivatives with respect to the spatial variable. The present paper seems to be the first one treating the unique solvability of these equations without imposing any {\em special\/} conditions on the relations between the coefficients or on their {\em derivatives\/}. This article in its spirit is similar to the author's recent articles \cite{KP}, \cite{Kr09_1}, \cite{Kr10_1}, and \cite{Kr10_2} and we spare the reader the common part of the comments about the literature, which can be found in the above references. The main idea, we use, originated from \cite{KP} and \cite{Kr09_1} and relies on application of special cut-off functions whose support evolves in time in a manner adapted to the drift terms. The paper consists of two parts: Sections \ref{section 11.11.1} trough \ref{section 1.1.2} are devoted to some general issues of the theory of SPDEs with growing coefficients and in Sections \ref{section 11.8.1} through \ref{section 1.1.4} we apply the results of the previous sections to show that the filtering equations corresponding to the Kalman-Bucy filter fall into the general theory. In a sense the methods of the first part of the present article arose as a combination of the methods from \cite{Kr10_1} and \cite{Kr10_2} which allow us to combine the method used for PDE equations with irregular (VMO) higher-order coefficients, growing lower-order coefficients, and $p>1$ with the methods which work in similar situation for SPDEs if $p=2$. Since we are interested in higher regularity of solutions (see, for instance, Theorem \ref{theorem 1.12.1}) we use the power of summability $p\geq2$ and, in contrast with \cite{Kr10_1}, this forces us to require some regularity of the higher-order coefficients. Roughly speaking we need the second-order coefficients of the deterministic part of the equation belong to VMO in $x$ and the first-order coefficients of the stochastic part to be uniformly continuous in $x$. In particular, the results of the present article do not generalize those of \cite{Kr10_1}. On the other hand, if we drop all stochastic terms, then we obtain the results of \cite{Kr10_2} for $p\geq2$, which by duality, available for deterministic equations, allows one to extend the result to full range $p>1$. Concerning the {\em deterministic\/} equations with growing coefficients in spaces with or without weights it is worth mentioning that (i) Equations in spaces with weights are treated, for instance, in \cite{CV1}, \cite{ChG}, \cite{FL}, \cite{Lu}, and \cite{MP1} for time independent coefficients, part of the result of which are extended in \cite{GL} to time-dependent Ornstein-Uhlenbeck operators; (ii) Equations in spaces without weights are treated, for instance, in \cite{LV}, \cite{MP}, \cite{PR}, and \cite{CF}. Some conclusions in the above cited papers are quite similar to ours but the corresponding assumptions are not as general in what concerns the regularity of the coefficients. However, these papers contain a lot of additional important information, which is probably impossible to obtain by using our methods. The second part of the article is devoted to the Kalman-Bucy filter. One can say that one of the sources of interest in SPDEs with growing coefficients is Zakai's equation for filtering density in the case of partially observable diffusion processes. This equation has divergence form which makes it possible to use the results of the first part of the article. In a very particular case of Gaussian processes the filtering density is given by the Kalman-Bucy filter. Generally, part of the coefficients of filtering equations in case of Gaussian processes grow. When the coefficients of an SPDE grow, it is quite natural to consider the equations in function spaces with weights which would restrict the set of solutions in such a way that all terms in the equation will be from the same space as the free terms. There are very many articles which use this idea in $\cL_{2}$- and $\cL_{p}$-settings (see, for instance, \cite{CV}, \cite{GK}, \cite{Gy93}, \cite{Gy97} and the references therein). Unfortunately, the application of the spaces with weights do not allow one to treat filtering equations corresponding to the Kalman-Bucy filter even without the so-called cross terms when the operators $\Lambda^{k}_{t}$ in \eqref{10.20.2} are of zeroth order. The main obstacle here is that the zeroth order coefficient of $\Lambda^{k}_{t}$ is a linear function of $x$. In the general theory, which we develop in this article, we do not allow it to grow either and we use an auxiliary function to ``kill'' this coefficient. The construction of this auxiliary function exploits a specific structure of the equation and allows us to transform the general filtering equation \eqref{10.20.2} to its ``reduced" form \eqref{9.29.01}, which does not contain the zeroth order term in the stochastic part. After that one can use a simple change of the unknown function shifting the $x$ variables in such a way that the stochastic part of \eqref{9.29.01} will disappear altogether and the equation will become a parabolic equation with time inhomogeneous and random Ornstein-Uhlenbeck operator. The fact that the operator is time inhomogeneous makes it impossible to apply any results based, for instance, on the semigroup approach and even specifically aimed at the Ornstein-Uhlenbeck operator, which one can find in the above mentioned recent articles such as \cite{ChG}, \cite{FL}, \cite{MP1}, or other results on elliptic operators with unbounded coefficients such as in \cite{PR}. The results of \cite{CV} are not applicable either because in \cite{CV} the zeroth-order coefficient is assumed to grow quadratically if the firs-order coefficients grow linearly. However, the results of \cite{GK} on general SPDEs with growing coefficients are applicable to the reduced form of the SPDE for the Kalman-Bucy filter and they provide existence and uniqueness theorems in Sobolev spaces with $p=2$ and weights depending on $t,x$ {\em and\/} $\omega$. By the way, a drawback of using weights depending on $t$ is that one cannot extract from the results for general SPDEs any result for deterministic elliptic equations. If one concentrates on $p=2$, then one can use the results from \cite{GL} where the Ornstein-Uhlenbeck time inhomogeneous operators are investigated in Sobolev spaces with Gaussian time dependent weight. Again this would allow one to investigate \eqref{9.29.01} in Sobolev spaces with $p=2$ and weights depending on $t,x$ and $\omega$. We deal with any $p\geq2$ and do not use weights. The article is organized as follows. In Section \ref{section 11.11.1} we introduce basic notation, function spaces, and equations. Section \ref{section 2.15.1} contains our main results concerning SPDEs. Section \ref{section 1.1.1} contains the proof of Theorem \ref{theorem 3.11.1} concerning an apriori estimate and Theorem \ref{theorem 1.12.1} about regularity properties of solutions. In Section \ref{section 6.9.5} we prove the existence Theorem \ref{theorem 3.16.1}. In Section \ref{section 1.1.2} we prove a version of It\^o's formula which allows us to use the results of the previous sections to derive the filtering equation without using anything from the filtering theory itself. We do it by following \cite{KZ} and \cite{Kr_10}. In Section \ref{section 11.8.1} we state our main result about the equation corresponding to Kalman-Bucy filter. We consider the so-called conditionally Gaussian process in the spirit of \cite{LS}. However, in contrast with \cite{LS}, our coefficients depend only on the current state of the two-component process under consideration and are not allowed to depend on the whole past of the observable component. In Section \ref{section 1.1.3} we consider the ``reduced'' form \eqref{9.29.01} of the main filtering equation \eqref{10.20.2}. The results of the previous sections turn out to be applicable to \eqref{9.29.01}. In the final Section \ref{section 1.1.4} we finish proving Theorems \ref{theorem 11.8.1} and \ref{theorem 11.11.1}, part of assertions of the former being proved in Section~\ref{section 1.1.3}. \mysection{General setting} \label{section 11.11.1} Let $(\Omega,\cF,P)$ be a complete probability space with an increasing filtration $\{\cF_{t},t\geq0\}$ of complete with respect to $(\cF,P)$ $\sigma$-fields $\cF_{t}\subset\cF$. Denote by $\cP=\cP(\{\cF_{t}\})$ the predictable $\sigma$-field in $\Omega\times(0,\infty)$ associated with $\{\cF_{t}\}$. Let $w^{k}_{t}$, $k=1,2,...$, be independent one-dimensional Wiener processes with respect to $\{\cF_{t}\}$. Let $\tau$ be a stopping time. We consider the second-order operator $L_{t}$ \begin{equation} \label{lu} L_{t} u_{t}(x) =D_{i}\big( a^{ij}_{t}( x)D_{j}u_{t}(x) +\gb^{i}_{t}(x)u_{t}(x)\big) + b^{i} _{t}(x) D_{i} u_{t}(x) -c _{t}(x) u_{t}(x), \end{equation} and the first-order operators $$ \Lambda^{k}_{t} u_{t}(x)=\sigma^{ik}_{t}(x)D_{i}u_{t}(x) +\nu^{k}_{t}(x)u_{t}(x) $$ acting on functions $u_{t}(x)$ defined on $\Omega\times\bR^{d+1}_{+}$, where $\bR^{d+1}_{+}= [0, \infty) \times \bR^d$, and given for $k=1,2,...$ (the summation convention is enforced throughout the article), where $$ D_{i}=\frac{\partial}{\partial x^{i}}. $$ We set $\bR_{+}=[0,\infty)$. Our main concern in the first part of the paper is proving the unique solvability of the equation \begin{equation} \label{2.6.4} du_{t}=(L_{t}u_{t}-\lambda u_{t}+D_{i}f^{i}_{t}+f^{0}_{t})\,dt +(\Lambda^{k}_{t}u_{t}+g^{k}_{t})\,dw^{k}_{t}, \quad t\leq\tau, \end{equation} with an appropriate initial condition at $t=0$, where $\lambda\geq0$ is a constant. The precise assumptions on the coefficients, free terms, and initial data will be given later. First we introduce appropriate function spaces. Fix a number $$ p\geq2, $$ and denote $\cL_{p}=\cL_{p}(\bR^{d})$. We use the same notation $\cL_{p}$ for vector- and matrix-valued or else $\ell_{2}$-valued functions such as $g_{t}=(g^{k}_{t})$ in \eqref{2.6.4}. For instance, if $u(x)=(u^{1}(x),u^{2}(x),...)$ is an $\ell_{2}$-valued measurable function on $\bR^{d}$, then $$ \|u\|^{p}_{\cL_{p}}=\int_{\bR^{d}}|u(x)|_{\ell_{2}}^{p} \,dx =\int_{\bR^{d}}\big( \sum_{k=1}^{\infty}|u^{k}(x)|^{2}\big)^{p/2} \,dx. $$ As usual, $$ W^{1}_{p}=\{u\in \cL_{p}: Du\in \cL_{p}\}, \quad \|u\|_{W^{1}_{p}}= \|u\|_{\cL_{p}}+\|Du\|_{\cL_{p}}, $$ where by $Du$ we mean the gradient with respect to $x$ of a function $u$ on $\bR^{d}$. Recall that $\tau$ is a stopping time and introduce $$ \bL _{p}(\tau):=\bL _{p}(\{\cF_{t}\},\tau):=\cL_{p}(\opar 0,\tau\cbrk,\cP, \cL_{p}), $$ $$ \bW^{1}_{p}(\tau):=\bW^{1}_{p}(\{\cF_{t}\},\tau):= \cL_{p}(\opar 0,\tau\cbrk,\cP, W^{1}_{p}), $$ $$ \bL_{p}=\bL_{p}(\infty), \quad\bW^{1}_{p}=\bW^{1}_{p}(\infty). $$ Remember that the elements of $\bL_{p}( \tau)$ need only belong to $\cL_{p}$ on a predictable subset of $\opar 0,\tau\cbrk$ of full measure. For the sake of convenience we will always assume that they are defined everywhere on $\opar 0,\tau\cbrk$ at least as generalized functions. Similar situation occurs in the case of $\bW^{1}_{p}( \tau)$. The following definition is most appropriate for investigating our equations if the coefficients of $L_{t}$ and $\Lambda^{k}_{t}$ are bounded. \begin{definition} \label{definition 3.16.1} Introduce $\cW^{1}_{p}(\tau)$, as the space of functions $u_{t} =u_{t}(\omega,\cdot)$ on $\{(\omega,t): 0\leq t\leq\tau(\omega),t<\infty\}$ with values in the space of generalized functions on $\bR^{d}$ and having the following properties: (i) We have $u_{0}\in \cL_{p}(\Omega,\cF_{0},\cL_{p})$; (ii) We have $u \in \bW^{1}_{p}(\tau )$; (iii) There exist $f^{i}\in \bL_{p}(\tau)$, $i=0,...,d$, and $g=(g^{1},g^{2},...)\in \bL_{p}(\tau)$ such that for any $\varphi\in C^{\infty}_{0}=C^{\infty}_{0}(\bR^{d})$ with probability 1 for all $t\in[0,\infty)$ we have $$ (u_{t\wedge\tau},\varphi)=(u_{0},\varphi) +\sum_{k=1}^{\infty}\int_{0}^{t}I_{s\leq\tau} (g^{k}_{s},\varphi)\,dw^{k}_{s} $$ \begin{equation} \label{1.2.1} + \int_{0}^{t}I_{s\leq\tau}\big((f^{0}_{s},\varphi)-(f^{i}_{s},D_{i}\varphi) \big)\,ds. \end{equation} In particular, for any $\phi\in C^{\infty}_{0}$, the process $(u_{t\wedge\tau},\phi)$ is $\cF_{t}$-adapted and (a.s.) continuous. In case that property (iii) holds, we write $$ du_{t}=(D_{i}f^{i}_{t}+f^{0}_{t})\,dt+g^{k}_{t}\,dw^{k}_{t}, \quad t\leq\tau. $$ Finally, set $\cW^{1}_{p}=\cW^{1}_{p}(\infty)$. \end{definition} \begin{remark} \label{remark 1.12.3} The reader understands that if $u$ is a generalized function on $\bR^{d}$, then $(u,\phi)$ represents the result of the action of $u$ on the test function $\phi\in C^{\infty}_{0}$. When $u$ is a locally integrable function, $(u,\phi)$ is the integral of the product $u\phi$. According to these notation $$ (f^{0}_{s},\varphi)-(f^{i}_{s},D_{i}\varphi)= (\bar{f}_{s},\phi), $$ where the function $\bar{f}_{s}$ with values in the space of generalized functions is defined by $\bar{f}_{s} =D_{i}f^{i}_{s}+f^{0}_{s}$. In the framework of Definition \ref{definition 3.16.1} we have $\bar{f} \in\cL_{p}(\opar0,\tau\cbrk,\cP,H^{-1}_{p})$, where $H^{-1}_{p}=(1-\Delta)^{1/2}\cL_{p}$. One also knows that any $\bar{f} \in\cL_{p}(\opar0,\tau\cbrk,\cP,H^{-1}_{p})$ is written as $\bar{f}_{s} =D_{i}f^{i}_{s}+f^{0}_{s}$ with some $f^{j}\in\bL_{p}(\tau)$. \end{remark} Also introduce the spaces of initial data in the same way as in \cite{Kr09}. \begin{definition} \label{definition 8.10.1} Let $u_{0}$ be an $\cF_{0}$-measurable function on $\Omega$ with values in the space of generalized functions on $\bR^{d}$. We write $u_{0}\in \text{\rm tr}\, \cW^{1}_{p}=\text{\rm tr}\,\cW^{1}_{p}(\cF_{0})$ if there exists a function $v \in \cW^{1}_{p} $ such that $d v_{t}=(\Delta v_{t}-v_{t})\,dt$, $t\in\bR_{+}$, and $v_{0}=u_{0}$. In such a case we set $$ \|u_{0}\|^{p}_{\text{\rm tr}\, \cW^{1}_{p}}=E\|v\|^{p}_{\bW^{1}_{p} }. $$ \end{definition} One knows that $\text{\rm tr}\, \cW^{1}_{p}$ is a Banach space, $v$ in the above definition is unique and $\cF_{0}$-measurable. We give the definition of solution of \eqref{2.6.4} adopted throughout the article and which in case the coefficients of $L_{t}$ and $\Lambda^{k}_{t}$ are bounded coincides with the one obtained by applying Definition \ref{definition 3.16.1}. \begin{definition} \label{definition 3.20.01} Let $f^{j}\in\bL_{p}(\tau)$, $j=0,...,d$, $g=(g^{1},g^{2},...) \in\bL_{p}(\tau)$. By a solution of \eqref{2.6.4} (relative to $\{\cF_{t}\}$) with initial condition $u_{0}\in \text{\rm tr}\, \cW^{1}_{p}$ we mean a function $u\in\bW^{1}_{p}(\tau)$ (not $\cW^{1}_{p}(\tau)$) such that (i) For any $\phi\in C^{\infty}_{0} $ the integrals in $$ (u_{t\wedge\tau},\phi)=(u_{0},\phi) +\sum_{k=1}^{\infty}\int_{0}^{t}I_{s\leq\tau} (\sigma^{ik}_{s}D_{i}u_{s}+\nu^{k}_{s}u_{s} +g^{k}_{s},\phi)\,dw^{k}_{s} $$ \begin{equation} \label{3.16.7} + \int_{0}^{t}I_{s\leq\tau}\big[(b^{i}_{s}D_{i}u_{s} -(c_{s}+\lambda)u_{s}+f^{0}_{s},\phi) -(a^{ij}_{s}D_{j}u_{s}+\gb^{i}_{s}u_{s}+ f^{i}_{s},D_{i}\phi) \big]\,ds \end{equation} are well defined and are finite for all finite $t\in\bR_{+}$ and the series converges uniformly on finite subinterval of $\bR_{+}$ in probability; (ii) For any $\phi\in C^{\infty}_{0} $ with probability one equation \eqref{3.16.7} holds for all $t\in\bR_{+}$. \end{definition} Observe that for any solution of \eqref{2.6.4} in the sense of the above definition and any $\phi\in C^{\infty}_{0}$ the process $(u_{t\wedge\tau},\phi)$ is continuous (a.s.) and $\cF_{t}$-adapted. Also notice that, if the coefficients of $L$ and $\Lambda^{k}$ are bounded, then any $u\in\cW^{1}_{p}(\tau)$ is a solution of \eqref{2.6.4} with appropriate free terms since if \eqref{1.2.1} holds, then \eqref{2.6.4} holds (always in the sense of Definition \ref{definition 3.20.01}) as well with $$ f^{i}_{t}-a^{ij}_{t}D_{j}u_{t}-\gb^{i}u_{t}, \quad i=1,...,d,\quad f^{0}_{t}+(c_{t}+\lambda)u_{t}-b^{i}_{t}D_{i}u_{t}, $$ $$ g^{k}_{t}-\sigma^{ik}D_{i}u_{t}-\nu^{k}_{t}u_{t} $$ in place of $f^{i}_{t}$, $i=1,...,d$, $f^{0}_{t}$, and $g^{k}_{t}$, respectively. \mysection{Main results for SPDEs} \label{section 2.15.1} For $\rho>0$ denote $B_{\rho}(x)=\{y\in\bR^{d}:|x-y|<\rho\}$, $B_{\rho}=B_{\rho}(0)$. \begin{assumption} \label{assumption 2.7.2} (i) The functions $a^{ij}_{t}(x)$, $\gb^{i}_{t}(x)$, $b^{i}_{t}(x)$, $c_{t}(x)$, $\sigma^{ik}_{t}(x)$, $\nu^{k}_{t}(x)$ are real valued, measurable with respect to $\cF\otimes\cB(\bR^{d+1}_{+})$, $\cF_{t}$-adapted for any $x$, and $c\geq 0$. (ii) There exists a constant $ \delta>0$ such that for all values of arguments and $\xi\in\bR^{d}$ $$ (a^{ij} - \alpha^{ij} ) \xi^{i} \xi^{j}\geq\delta|\xi|^{2},\quad |a^{ij}|\leq \delta^{-1} , \quad |\nu|_{\ell_{2}}\leq \delta^{-1}, $$ where $\alpha^{ij} =(1/2)(\sigma^{i\cdot},\sigma^{j\cdot}) _{\ell_{2}}$. Also, the constant $\lambda\geq0$. (iii) For any $x\in \bR^{d}$ (and $\omega$) the function \begin{equation} \label{9.3.01} \int_{B_{1}}(|\gb_{t}(x+y)|+|b_{t}(x+y)|+|c_{t}(x+y)|)\,dy \end{equation} is locally integrable to the $p'$th power on $\bR_{+}=[0,\infty)$, where $p'=p/(p-1)$. \end{assumption} Notice that the matrix $a=(a^{ij})$ need not be symmetric. Also notice that in Assumption \ref{assumption 2.7.2} (iii) the ball $B_{1}$ can be replaced with any other ball without changing the set of admissible coefficients $\gb,b,c$. Recall that as is well known if $u\in\bW^{1}_{p}(\tau)$, then owing to the boundedness of $\nu$ and $\sigma$ and the fact that $Du,u,g\in\bL_{p}(\tau)$, $p\geq2$, the first series on the right in \eqref{3.16.7} converges uniformly in probability and the series is a continuous local martingale. Furthermore, if we denote it by $m_{t}$, then for any $T\in\bR_{+}$ $$ E\sup_{t\leq T}|m_{t}|^{p}\leq N E\big(\sum_{k=1}^{\infty}\int_{0}^{\tau\wedge T} (\sigma^{ik}_{s}D_{i}u_{s}+\nu^{k}_{s}u_{s} +g^{k}_{s},\phi)^{2}\,ds\big)^{p/2} $$ $$ \leq N\|\phi\| _{\cL_{1}}^{p/2} E\big( \int_{0}^{\tau\wedge T}\sum_{k=1}^{\infty} (|\sigma_{s}^{ik}|^{2}|D_{i}u_{s}|^{2}+|\nu^{k}_{s}|^{2} |u_{s}|^{2} +|g^{k} _{s}|^{2},|\phi|) \,ds\big)^{p/2} $$ \begin{equation} \label{9.3.3} \leq N(\|u\|^{p}_{\bW^{1}_{p}(\tau)}+ \|g\|^{p}_{\bL_{p}(\tau)}), \end{equation} where the constants $N$ depend only on $\phi$, $d$, $p$, $\delta$, and $T$. \begin{assumption} \label{assumption 2.4.1} There exists a function $\kappa(r)$, $r\in\bR_{+}$, such that $\kappa(0+)=0$ and for any $\omega\in\Omega$, $t\geq0$, $x,y\in\bR^{d}$, and $i =1,...,d$ we have $$ |\sigma_{t}^{i\cdot}(x)-\sigma^{i\cdot}_{t}(y)|_{\ell_{2}} \leq\kappa(|x-y|). $$ \end{assumption} The following assumptions contain parameters $ \gamma_{a},\gamma_{b} \in(0,1]$, whose values will be specified later. They also contain constants $K\geq0$, $\rho_{0}, \rho_{1} \in(0,1]$ which are fixed. \begin{assumption} \label{assumption 2.4.01} For any $\omega\in\Omega$, $\rho\in(0,\rho_{0}]$, $t\geq0$, and $i,j=1,...,d$ we have \begin{equation} \label{2.9.5} \rho^{-2d-2}\int_{t}^{t+\rho^{2}}\bigg( \sup_{x\in \bR^{d}}\int_{B_{\rho}(x)}\int_{B_{\rho}(x)} |a^{ij}_{s}( y)-a^{ij}_{s}( z)|\,dydz\bigg) \,ds\leq\gamma_{a}. \end{equation} \end{assumption} Obviously, the left-hand side of \eqref{2.9.5} is less than $$ N(d)\sup_{t\geq0}\sup_{|x-y|\leq2\rho}|a^{ij}_{t}( x ) -a^{ij}_{t}(y)|, $$ which implies that Assumption \ref{assumption 2.4.01} is satisfied with any $\gamma_{a} >0$ if, for instance, $a$ is uniformly continuous in $x$ uniformly in $\omega$ and $t$. Recall that if $a$ is independent of $t$ and for any $\gamma_{a} >0$ there is a $\rho_{0}>0$ such that Assumption \ref{assumption 2.4.01} is satisfied, then one says that $a$ is in VMO. We take and fix a number $q=q(d,p) $ such that \begin{equation} \label{8.11.3} q\geq\max(d,p)\quad\hbox{\rm if}\quad p\ne d, \quad q>d\quad\hbox{\rm if}\quad p= d. \end{equation} \begin{assumption} \label{assumption 3.11.1} For any $\omega\in\Omega$, $ \gb:=(\gb^{1} ,...,\gb^{d} )$, $ b:=(b^{1} ,...,b^{d} )$, and $(t,x)\in\bR^{d+1}$ we have $$ \int_{B_{\rho_{1}}(x)}\int_{B_{\rho_{1}}(x)}|\gb_{t}( y) -\gb_{t}( z)|^{q }\,dydz + \int_{B_{\rho_{1}}(x)}\int_{B_{\rho_{1}}(x)}|b_{t}( y) -b_{t}( z)|^{q}\,dydz $$ $$ + \int_{B_{\rho_{1}}(x)}\int_{B_{\rho_{1}}(x)}|c_{t}( y) -c_{t}( z)|^{q}\,dydz \leq KI_{q>d}+\rho_{1}^{d}\gamma_{b}. $$ \end{assumption} Obviously, Assumption \ref{assumption 3.11.1} is satisfied if $b$, $\gb$, and $c$ are independent of $x$. They also are satisfied with any $q>d$, $\gamma_{b}=0$, and $\rho_{1}=1$ on the account of choosing $K$ appropriately if, say, $$ |\gb_{t}( x ) -\gb_{t}(y)|+|b_{t}( x ) -b_{t}( y)|+|c_{t}( x ) -c_{t}( y)|\leq K_{1} $$ whenever $|x-y|\leq1$, where $K_{1}$ is a constant. In particular, Assumption \ref{assumption 3.11.1} is satisfied if $\gb$, $b$, and $c$ are globally Lipschitz continuous: $$ |\gb_{t}( x ) -\gb_{t}(y)|+|b_{t}( x ) -b_{t}( y)| $$ \begin{equation} \label{1.15.1} +|c_{t}( x ) -c_{t}( y)|\leq K_{1}|x-y|,\quad\forall x,y\in\bR^{d},t\geq0. \end{equation} We see that Assumption \ref{assumption 3.11.1} allows $b$, $\gb$, and $c$ growing linearly in $x$. Here is our result on apriori estimates of solutions of \eqref{2.6.4}. \begin{theorem} \label{theorem 3.11.1} There exist $$ \gamma_{a} =\gamma_{a}(d,\delta,p),\quad \gamma_{b} =\gamma_{b}(d,\delta,p,\kappa,\rho_{0} )\in(0,1], $$ $$ N=N(d,\delta,p,\kappa,\rho_{0} ), \quad \lambda_{0}=\lambda_{0}(d,\delta, p,\kappa, \rho_{0},\rho_{1},K)\geq1 $$ such that, if the above assumptions are satisfied and $\lambda\geq \lambda_{0}$ and $u $ is a solution of \eqref{2.6.4} with initial data $u_{0}\in \text{\rm tr}\, \cW^{1}_{p}$ and some $f^{j},g \in\bL_{p}(\tau)$, then $$ \lambda\|u\|^{2}_{\bL _{p}(\tau)}+\|Du\|^{2}_{\bL _{p}(\tau)} \leq N\big(\sum_{i=1}^{d}\|f^{i}\|^{2}_{\bL _{p}(\tau)} +\|g\|^{2}_{\bL _{p}(\tau)} \big) $$ \begin{equation} \label{3.11.2} +N\lambda^{-1}\|f^{0}\|^{2}_{\bL _{p}(\tau)} +N \|u_{0}\|^{2}_{\text{\rm tr}\, \cW^{1}_{p}} . \end{equation} \end{theorem} \begin{remark} \label{remark 9.27.1} There is an unusual property of $u_{t}$, which is nontrivial even if $ f^{j}_{t} =g^{k}_{t} \equiv0$. Namely, assume that $g\equiv0$. Take a predictable $\ell_{2}$-valued process $\xi_{t}$ such that $(\nu_{t},\xi_{t})_{\ell_{2}}\geq0$ and $(\nu_{t},\xi_{t})_{\ell_{2}}$ and $(\sigma^{i\cdot}_{t}, \xi_{t})$ are independent of $x$ (which happens, for instance, if $\nu=0$ and $\sigma$ is independent of $x$) and $$ \int_{0}^{\tau}|\xi_{t}|_{\ell_{2}}^{2}\,dt<\infty $$ (a.s.) and assume that $E\rho_{\tau}(\xi)=1$, where $$ \rho_{t}(\xi)=\rho_{t}(\xi,dw):= \exp\big(-\int_{0}^{t}\xi^{k}_{s}\,dw^{k}_{s} -\tfrac{1}{2}\int_{0}^{t}|\xi_{s}|_{\ell_{2}}^{2}\,ds\big) . $$ Then the assertion of Theorem \ref{theorem 3.11.1} holds with the {\em same\/} $\gamma_{a}$, $\gamma_{b}$, $\lambda_{0}$, and $N$ if we understand $\|v\|_{\bL_{p}(\tau)}^{p}$ for all $v$'s as $$ E\rho_{\tau}\int_{0}^{\tau}\|v_{t}\|^{p}_{\cL_{p}}\,dt. $$ Indeed, one can change the probability measure by using Girsanov's theorem. This will add a new drift term in the deterministic part of \eqref{2.6.4} and this additional drift depends only on $(\omega, t)$. This will also add the term $-(\nu_{t},\xi_{t})_{\ell_{2}}u_{t}\,dt$, where $(\nu_{t},\xi_{t})_{\ell_{2}}$ is nonnegative and also independent of $x$. Then the result follows immediately from Theorem~\ref{theorem 3.16.1}. \end{remark} Theorem \ref{theorem 3.11.1} admits the following version if $\tau$ is bounded. \begin{theorem} \label{theorem 1.14.1} Let $T\in(0,\infty)$ be a constant and suppose that $\tau\leq T$. Assume that the above assumptions are satisfied with $\gamma_{a}$ and $\gamma_{b}$ from Theorem~\ref{theorem 3.11.1}. Let $\lambda=0$ and let $u $ be a solution of \eqref{2.6.4} with initial data $u_{0}\in \text{\rm tr}\, \cW^{1}_{p}$ and some $f^{j},g \in\bL_{p}(\tau)$. Then \begin{equation} \label{1.14.2} \|u\|^{2}_{\bW^{1} _{p}(\tau)} \leq N\big(\sum_{i=0}^{d}\|f^{i}\|^{2}_{\bL _{p}(\tau)} +\|g\|^{2}_{\bL _{p}(\tau)} +\|u_{0}\|^{2}_{\text{\rm tr}\, \cW^{1}_{p}}\big) , \end{equation} where $N=N(d,\delta, p,\kappa, \rho_{0},\rho_{1},K,T)$. \end{theorem} This result is a trivial consequence of Theorem \ref{theorem 3.11.1} since, for any constant $\mu$, the function $v_{t}:=u_{t}e^{-\mu t}$ satisfies \eqref{2.6.4} with $\lambda+\mu$, $f^{j}_{t}e^{-\mu t}$, and $g^{k}_{t}e^{-\mu t}$ in place of $\lambda$, $f^{j}_{t}$, and $g^{k}_{t}$, respectively. If $\mu$ is large enough and $\tau\leq T$, estimate \eqref{3.11.2} for $v$ implies \eqref{1.14.2} indeed. \begin{remark} Theorems \ref{theorem 3.11.1} and \ref{theorem 1.14.1} provide uniqueness of solutions of \eqref{2.6.4}. The apriori estimates \eqref{3.11.2} and \eqref{1.14.2} can also be used to investigate continuous dependence of solutions on the coefficients and other data. \end{remark} To prove the existence we need stronger assumptions because, generally, Assumption \ref{assumption 3.11.1} does not guarantee that $$ D_{i}(\gb^{i}_{t}u_{t})+b^{i}_{t}D_{i}u_{t}-c_{t}u_{t} $$ can be written even locally as $D_{i}\hat{f}^{i}_{t}+\hat{f}^{0}_{t}$ with $\hat{f}^{j}\in\bL_{p}(\tau)$ if we only know that $u\in\bW^{1}_{p}(\tau)$ even if $\gb$, $b$, and $c$ are independent of $x$. We can only prove our Lemma \ref{lemma 3.23.2} if we have a certain control on this expression. \begin{assumption} \label{assumption 3.16.1} For any $x\in \bR^{d}$ (and $\omega$) the function \eqref{9.3.01} is locally integrable to the power $ p /(p-2)$ (locally bounded if $p=2$) on $\bR_{+}=[0,\infty)$. \end{assumption} \begin{remark} \label{remark 1.15.1} Assumptions \ref{assumption 3.11.1} and \ref{assumption 3.16.1} are both satisfied if the global Lipschitz condition \eqref{1.15.1} holds and $b_{t}(0)$, $\gb_{t}(0)$, and $c_{t}(0)$ are bounded for each~$\omega$. \end{remark} \begin{theorem} \label{theorem 3.16.1} Let the above assumptions be satisfied with $\gamma_{a} $ and $\gamma_{b}$ taken from Theorem \ref{theorem 3.11.1}. Take $\lambda\geq\lambda_{0}$, where $\lambda_{0}$ is defined in Theorem \ref{theorem 3.11.1}, and take $u_{0}\in\text{\rm tr}\, \cW^{1}_{p}$. Then there exists a unique solution of \eqref{2.6.4} with initial condition $u_{0}$. \end{theorem} \begin{remark} \label{remark 9.29.2} If the stopping time $\tau$ is bounded, then in the above theorem one can take $\lambda_{0}=0$. This is shown by the same argument as after Theorem~\ref{theorem 1.14.1}. \end{remark} In general the continuity properties in $t$ of the solution from Theorem \ref{theorem 3.16.1} are unknown. For instance, we do not know if $\|u_{t\wedge\tau}\phi\|_{\cL_{p}}$ is continuous (a.s) for any $\phi\in C^{\infty}_{0}$. However, under stronger assumptions we can say more about regularity of $u$. In the following theorem by $H^{\gamma}_{p}$ we mean $(1-\Delta)^{-\gamma/2}\cL_{p}$. \begin{theorem} \label{theorem 1.12.1} Under the above assumptions suppose that for each $x\in\bR^{d}$ the function \eqref{9.3.01} is {\em bounded\/} on $\opar0,\tau\cbrk$. Then the (unique) solution $u$ possesses the following properties: (i) For any $\phi\in C^{\infty}_{0}$ we have $\phi u\in\cW^{1}_{p}(\tau)$; (ii) For any $\phi\in C^{\infty}_{0}$ the process $u_{t\wedge\tau}\phi$ is continuous on $\bR_{+}$ as an $\cL_{p}$-valued process (a.s.); (iii) If $p>2$ and $\tau$ is bounded and we have two numbers $\alpha$ and $\beta$ such that $$ \frac{2}{p}<\alpha<\beta\leq1, $$ then for any $\phi\in C^{\infty}_{0}$ (a.s.) $$ u\phi\in C^{\alpha/2-1/p}([0,\tau], H^{1-\beta}_{p}). $$ In particular, if $p>d+2$, then (a) for any $\varepsilon \in(0,\varepsilon_{0}]$, with $$ \varepsilon_{0}=1-\frac{d+2}{ p},$$ (a.s.) for any $t\in[0,\tau]$ we have $u_{t}\phi\in C^{\varepsilon_{0}- \varepsilon}(\bR^{d})$ and the norm of $u_{t}\phi$ in this space is bounded as a function of $t$; (b) for any $\varepsilon$ as in (a) (a.s.) for any $x\in\bR^{d}$ we have $u_{\cdot}(x)\phi(x)\in C^{(\varepsilon_{0}- \varepsilon)/2}([0,\tau])$ and the norm of $ u_{\cdot}(x)\phi(x)$ in this space is bounded as a function of $x$. \end{theorem} Observe that assertions (ii) and (iii) of Theorem \ref{theorem 1.12.1} follow from assertion (i) proved in Remark \ref{remark 1.12.1}. In case of assertion (ii) this is shown in \cite{Kr09_3}. The main part of assertion (iii) follows from assertion (i) and Corollary 4.12 \cite{Kr01}. By applying Sobolev's embedding theorems assertion (iii) (a) is obtained after taking $\alpha$ and $\beta$ close to $2/p$ and (iii) (b) after taking $\alpha$ and $\beta$ close to $1-d/p$. \begin{remark} Let $p_{1},p_{2}\in[2,\infty)$, let $\tau$ be bounded (cf. Remark \ref{remark 9.29.2}), and let the assumptions of Theorem \ref{theorem 3.16.1} be satisfied for any $p\in[p_{1},p_{2}]$ with $\gamma_{a}$ and $\gamma_{b}$ which are suitable for all $p\in[p_{1},p_{2}]$. Then it turns out that the solution from Theorem \ref{theorem 3.16.1} corresponding to $p=p_{1}$ coincides with the one obtained for $p=p_{2}$. This fact is obtained in the same way as the proof of Theorem 3.4 of \cite{Kr10_2} is obtained from the proof of Theorem 3.3 \cite{Kr10_2}. \end{remark} Our last main result on general SPDEs bears on the measurability of $u_{t}$ with respect to $\sigma$-fields which are smaller than $\cF_{t}$. It will be used in Section \ref{section 1.1.3} and this is the reason why we use the somewhat strange notation $\tilde{y}_{t}$ and $\tilde{\sfB}^{k}_{t}$ below. We suppose that all the above assumptions are satisfied with $\gamma_{a}$ and $\gamma_{b}$ taken from Theorem \ref{theorem 3.11.1} and let $\tilde{\cF}_{t}$, $t\geq0$, be a filtration of complete with respect to $\cF,P$ $\sigma$-fields such that $ \cF _{t}\supset \tilde{\cF}_{t}$. Our aim is to show that sometimes $u_{t}$ is $\tilde{\cF}_{t}$-adapted even if some terms in \eqref{2.6.4} are not $\tilde{\cF}_{t}$-adapted. However, the equation is assumed to have a special structure. The result is not surprising because in the notation, introduced below, equation $$ du_{t}= (\Lambda^{k}_{t}u_{t}+g^{k}_{t})\,dw^{k}_{t} $$ \begin{equation} \label{10.6.2} +(L_{t}u_{t}+\hat{b}_{t}^{i}D_{i}u_{t}-\hat{c}_{t}u_{t} +D_{i}f^{i}_{t}+f^{0}_{t}+\hat{f}_{t})\,dt, \quad t\leq\tau \end{equation} is written as \begin{equation} \label{10.7.01} du_{t}= (\Lambda^{k}_{t}u_{t}+g^{k}_{t})\,d\tilde{y}^{k}_{t} +(L_{t}u_{t} +D_{i}f^{i}_{t}+f^{0}_{t} )\,dt, \quad t\leq\tau, \end{equation} \begin{theorem} \label{theorem 10.6.1} Fix a number $T\in(0,\infty)$. Assume that we are given an $\ell_{2}$-valued process $\tilde{\sfB}_{t}$ which is $ \cF _{t}$-adapted, jointly measurable with respect to $(\omega,t)$, and such that $|\tilde{\sfB}_{t}|_{\ell_{2}}$ is locally square integrable on $\bR_{+}$ and $E\rho_{T}=1$, where $$ \rho_{t} =\rho_{t}(\tilde{\sfB},dw) =\exp(-\int_{0}^{t} \tilde{\sfB}^{k}_{s}\,dw^{k}_{s} -\tfrac{1}{2}\int_{0}^{t}|\tilde{\sfB}_{s}|_{\ell_{2}}^{2}\,ds). $$ Suppose that Assumption \ref{assumption 2.7.2} (i) is satisfied with $\tilde{\cF}_{t}$ in place of $\cF_{t}$ and the processes $$ \tilde{y}_{t}^{k}=w_{t}^{k} +\int_{0}^{t}\tilde{\sfB}^{k}_{s} \,ds $$ are $\tilde{\cF}_{t}$ adapted. Introduce $$ \bar{b}^{i}_{t}(x)=\sigma^{ik}_{t}(x)\tilde{\sfB}^{k}_{t}, \quad \bar{c}_{t}(x)=-\nu^{k}_{t}(x)\tilde{\sfB}^{k}_{t} $$ and suppose that $b+\bar{b} $ and $c+\bar{c}$ satisfy Assumption \ref{assumption 3.11.1} with $\gamma_{b}$ from Theorem \ref{theorem 3.11.1}, for any $x\in \bR^{d}$ (and $\omega$) we have $\bar{c}_{t}(x)\leq K$, and the function \begin{equation} \label{9.3.1} \int_{B_{1}}(|\bar{b}_{t}(x+y)| +|\bar{c}_{t}(x+y)|)\,dy \end{equation} is locally integrable to the power $ p /(p-2)$ (locally bounded if $p=2$) on $\bR_{+}$. Let $\tau$ be an $\tilde{\cF}_{t}$-stopping time such that $\tau\leq T$. Then, for any initial data $u_{0}\in \text{\rm tr}\, \cW^{1}_{p}(\tilde{\cF}_{0})$ and $f^{j},g \in\bL _{p}(\{\tilde{\cF}_{t}\},\tau)$ such that $\tilde{f}:=(g,\tilde{\sfB})_{\ell_{2}} \in\bL _{p}(\{ \cF _{t}\},\tau)$, (i) equation \eqref{10.6.2} has a unique solution $u$ relative to $\{ \cF _{t}\}$ in the sense of Definition \ref{definition 3.20.01}, (ii) for any $\phi\in C^{\infty}_{0}$ the process $(u_{t\wedge\tau},\phi)$ is $\tilde{\cF}_{t}$-adapted. \end{theorem} Proof. Owing to the argument after Theorem~\ref{theorem 1.14.1} allowing us to introduce as large $\lambda$ as we wish, assertion (i) follow immediately from Theorem \ref{theorem 3.16.1}. To prove (ii) we use a change of measure. Define $\tilde{P}(d\omega)=\rho_{T}(\omega)\,P(d\omega)$, notice that by Girsanov's theorem the processes $\tilde{y}^{k}_{t}$, $t\leq T$, are independent Wiener processes with respect to $\tilde{P}, \cF_{t}$. By assumption they are $\tilde{\cF}_{t}$-adapted and since $\tilde{\cF}_{t}\subset\cF_{t}$ the increments $\tilde{y}^{k}_{t+s} -\tilde{y}^{k}_{t}$ are independent of $\tilde{\cF}_{t}$ if $s\geq0$. Thus $(\tilde{y}^{k}_{t},\tilde{\cF}_{t})$ are independent Wiener processes. Introduce $\tilde{E}$ as the expectation sign relative to $\tilde{P}$. After rewriting \eqref{10.6.2} in form \eqref{10.7.01} and applying Theorems \ref{theorem 3.11.1} and \ref{theorem 3.16.1} we get that there exists a unique solution $\tilde{u}$ of \eqref{10.6.2} with initial data $u_{0}$ relative to $\{\tilde{\cF}_{t}\}$ in the sense of Definition \ref{definition 3.20.01} on the new probability space, that is with $\bL_{p}(\tau)$ and $\bW^{1}_{p}(\tau)$ replaced with $\tilde{\,\bL}_{p}(\{\tilde{\cF}_{t}\},\tau)$ and $\!\tilde{\bW}^{1}_{p}(\{\tilde{\cF}_{t}\},\tau)$, respectively, where the norms in these spaces are defined as $$ \tilde{E}\int_{0}^{\tau}\|u_{t}\|^{p}_{\cL_{p}}\,dt\quad \hbox{and}\quad \tilde{E}\int_{0}^{\tau}\|u_{t}\|^{p}_{W^{1}_{p}}\,dt $$ raised to the power $1/p$, respectively. Now for $n\geq2$ we introduce $\cF_{t}$-stopping times $$ \tau_{n}=\tau\wedge\inf\{t\geq0:\rho_{t}\leq 1/n\} $$ and observe that $$ E\int_{0}^{\tau_{n}}\|\tilde{u}_{t}\|^{p}_{\cL_{p}}\,dt \leq n E\rho_{\tau_{n}} \int_{0}^{\tau_{n}}\|\tilde{u}_{t}\|^{p}_{\cL_{p}}\,dt= n\tilde{E}\int_{0}^{\tau_{n}}\|\tilde{u}_{t}\|^{p}_{\cL_{p}}\,dt<\infty. $$ Similar estimates hold if we replace $\cL_{p}$ with $W^{1}_{p}$. By recalling that $\tilde{\cF_{t}}\subset\cF_{t}$, we conclude that $\tilde{u}$ is a solution of \eqref{10.6.2} relative to $\{\cF_{t}\}$ with $\tau_{n}$ in place of $\tau$. By uniqueness, in the sense of distributions $\tilde{u}_{t }I_{t\leq\tau_{n}}= u_{t }I_{t\leq\tau_{n}}$ for almost all $(\omega,t)$, that is, $(\tilde{u}_{t },\phi)I_{t\leq\tau_{n}}= (u_{t },\phi)I_{t\leq\tau_{n}}$ for almost all $(\omega,t)$ for each fixed $\phi \in C^{\infty}_{0}$. Then it follows from the integral form of \eqref{10.6.2} that for each $\phi \in C^{\infty}_{0}$ with probability one $(\tilde{u}_{t\wedge\tau_{n} },\phi) = (u_{t \wedge\tau_{n} },\phi) $ for all $t$. Upon letting $n\to\infty$ we replace $\tau_{n}$ with $\tau$ and it only remains to observe that $(\tilde{u}_{t\wedge\tau },\phi)$ is $\tilde{\cF}_{t}$-measurable. The theorem is proved. The following is almost identical to Remark 3.5 of \cite{Kr10_1}. \begin{remark} We do not use the spaces with weights. However, there is a trivial and since very long time known way how to use results like ours for treating equations in spaces with weights. For instance, let $\psi_{t}(x)>0$ be a nonrandom smooth function on $\bR^{d+1}$. Introduce, $\partial_{t} =\partial/\partial t$, $$ \hat{\gb}^{i}_{t}=\gb^{i}_{t}-a^{ij}_{t}D_{j}\ln\psi_{t}, \quad \hat{b}^{i}_{t}=b^{i}_{t}-a^{ij}_{t}D_{j}\ln \psi_{t}, $$ $$ \hat{c}_{t}=c_{t}+(b^{i}_{t}+\gb^{i}_{t})D_{i}\ln\psi_{t} -a^{ij}_{t}(D_{i}\ln\psi_{t})D_{j}\ln\psi_{t}-\partial_{t}\ln\psi_{t}, $$ $$ \hat{\nu}^{k}_{t}=\nu^{k}_{t}-\sigma^{ik}_{t}D_{i}\ln\psi_{t}, $$ $$ \hat{f}^{i}_{t}=\psi_{t} f^{i}_{t},\quad i=1,...,d,\quad \hat{f}^{0}_{t}=f^{0}_{t}\psi_{t}-f^{i}_{t}D_{i}\psi_{t},\quad \hat{g}^{k}_{t}=g^{k}_{t}\psi_{t}. $$ Suppose that, if we replace $b$, $\gb$, $c$, and $\nu$ with $\hat{b}$, $\hat{\gb}$, $\hat{c}$, and $\hat{\nu}$, respectively, then Assumptions \ref{assumption 2.7.2}, \ref{assumption 3.11.1}, and \ref{assumption 3.16.1} are satisfied with $\gamma_{a}$ and $\gamma_{b}$ from Theorem \ref{theorem 3.11.1}. Finally, assume that $\hat{f}^{j},\hat{g} \in\bL_{2}(\tau)$ and $u_{0}\psi_{0}\in\text{\rm tr}\,\cW^{1}_{p}$. Then it turns out that for $\lambda\geq\lambda_{0}$ ($\lambda_{0}$ is taken from Theorem \ref{theorem 3.11.1}) equation \eqref{2.6.4} has a unique solution $u$ such that $u\psi\in\bW^{1}_{p}(\tau)$. This fact is almost trivial since $u$ satisfies \eqref{2.6.4} if and only if $v:=u\psi$ satisfies the version of \eqref{2.6.4} which is obtained as the result of the replacements described above and also the replacement of $f^{j},g$ with $\hat{f}^{j},\hat{g}$, respectively. In addition, the natural estimate of the $\bW^{1}_{p}(\tau)$-norm of $v$ gives an estimate of $u$ in an appropriate space with weights. As a specification of the above, in the setting of Remark \ref{remark 1.15.1} take a $T\in(0,\infty)$, set $\tau=T$, and for $\theta\in(0,\infty)$ introduce $$ \ln \psi_{t}(x)=-\theta e^{\theta^{2}(t-T)}\sqrt{1+|x|^{2}}. $$ Obviously, $D_{i}\ln\psi$ are bounded for $t\leq T$. Furthermore, it is not hard to see that if $\theta$ is large enough, then $\hat{c}_{t}\geq0$ for $t\leq T$. Also, if $|x-y|\leq1$, then owing to the fact that $|D_{ij}\ln\psi_{t}(x)| \leq N(1+|x|)^{-1}$ for $t\leq T$, where $N$ is a constant, we have $$ |b^{i}_{t}(x)D_{i}\ln\psi_{t}(x)-b^{i}_{t}(y)D_{i}\ln\psi_{t}(y)| $$ $$ \leq |(b^{i}_{t}(x)-b^{i}_{t}(y))D_{i}\ln\psi_{t}(x)| +N(1+|x|)|D\ln\psi_{t}(x)- D\ln\psi_{t}(y)| $$ $$ \leq K|D\ln\psi_{t}(x)|+N $$ for $t\leq T$. Estimates similar to this one show that $\hat{b}$, $\hat{\gb}$, and $\hat{c}$ satisfy Assumption \ref{assumption 3.11.1} for $t\leq T$. By what is said in the beginning of the current remark, if $u_{0}\psi_{0}\in \text{\rm tr}\,\cW^{1}_{p}$ (for instance, $u_{0}(x)=x^{1}$), then \eqref{2.6.4} has a unique solution $u$ such that $u\psi\in \bW^{1}_{p}(T)$. Since $D\ln\psi$ is bounded, the inclusion $u\psi\in \bW^{1}_{p}(T)$ is equivalent to $u\psi\in \bL_{p}(T)$, $\psi Du\in \bL_{p}(T)$. To the best of the author's knowledge even in this special case the result in this generality was not known before. \end{remark} \mysection{Proof of Theorems \protect\ref{theorem 3.11.1} and \protect\ref{theorem 1.12.1}} \label{section 1.1.1} In this section we suppose that Assumptions \ref{assumption 2.7.2}, \ref{assumption 2.4.1}, \ref{assumption 2.4.01}, and \ref{assumption 3.11.1} are satisfied with some $\gamma_{a}, \gamma_{b}\in(0,1]$ and start by showing that the requirement (i) of Definition \ref{definition 3.20.01} is automatically satisfied for any $u\in\bW^{1}_{p}(\tau)$. Take a nonnegative $ \xi\in C^{\infty}_{0}(B_{\rho_{1}})$ with unit integral and define $$ \bar{b}_{s}(x)=\int_{B_{\rho_{1}}}\xi(y) b_{s}(x-y) \,dy,\quad \bar{\gb}_{s}(x) =\int_{B_{\rho_{1}}}\xi(y) \gb_{s}(x-y) \,dy, $$ \begin{equation} \label{6.28.3} \bar{c}_{s}(x)=\int_{B_{\rho_{1}}}\xi(y) c_{s}(x-y) \,dy. \end{equation} We may assume that $|\xi|\leq N(d)\rho_{1}^{-d}$. \begin{remark} \label{remark 1.12.1} By Corollary 5.4 of \cite{Kr10_2} , for $x_{0}\in\bR^{d}$, $v\in\cL_{p}$, $\phi\in W^{1}_{p'}$, and $u\in W^{1}_{p}$ we have $$ (|b _{s}-\bar{b} _{s}(x_{0})|I_{B_{\rho_{1}}(x_{0})} v,|\phi|) \leq N \| v\| _{\cL_{p}} \|\phi\|_{W^{1}_{p'}} , $$ \begin{equation} \label{1.12.1} \|I_{B_{\rho_{1}}(x_{0})} |\gb_{s}-\bar{\gb}_{s}(x_{0})|\,u \|_{\cL_{p}} +\|I_{B_{\rho_{1}}(x_{0})}|c_{s}-\bar{c}_{s}(x_{0})|\,u\|_{\cL_{p}} \leq N\|u\|_{W^{1}_{p}}, \end{equation} where $N=N(d,p,\rho_{1},K) $. In particular, \begin{equation} \label{1.12.2} (|b _{s}|I_{B_{\rho_{1}}(x_{0})} v,|\phi|) \leq (N+|\bar{b} _{s}(x_{0})|) \| v\| _{\cL_{p}} \|\phi\|_{W^{1}_{p'}} , \end{equation} the latter implying that $|b _{s}|I_{B_{\rho_{1}}(x_{0})} v\in H^{-1}_{p}$. It is also seen that if $u\in\bW^{1}_{p}(\tau)$ and $|\bar{b} _{s}(x_{0})|$ is a {\em bounded\/} function on $\opar0,\tau\cbrk$, then $$ I_{B_{\rho_{1}}}(x_{0})b^{i} D_{i}u \in\cL_{p}(\opar0,\tau\cbrk,\cP,H^{-1}_{p}). $$ Similarly, \begin{equation} \label{1.13.1} \|I_{B_{\rho_{1}}(x_{0})} |\gb_{s}|\,u \|_{\cL_{p}} +\|I_{B_{\rho_{1}}(x_{0})}|c_{s}|\,u\|_{\cL_{p}} \leq (N+|\bar{\gb}_{s}(x_{0})|+ |\bar{c}_{s}(x_{0})|)\|u\|_{W^{1}_{p}}. \end{equation} By the way, Remark \ref{remark 1.12.3} now shows that under the conditions of Theorem \ref{theorem 1.12.1} for any solution $u$ of \eqref{2.6.4} and $\phi\in C^{\infty}_{0}$ with support lying in a ball of radius $\rho_{1}$ we have $u\phi\in\cW^{1}_{p}(\tau)$. Of course, the restriction on the size of support of $\phi$ is easily removed and this proves assertion (i) of Theorem \ref{theorem 1.12.1}. \end{remark} \begin{lemma} \label{lemma 6.27.1} Let $R\in(0,\infty)$. Then there exists a sequence of bounded stopping times $\tau_{n}\to\infty$ such that for any $\omega\in\Omega$, $u\in\cL_{p}((0,\tau),W^{1}_{p})$, and $\phi\in C^{\infty}_{0}(B_{R})$ $$ \int_{0}^{\tau_{n}\wedge\tau} (|(b^{i}_{s}D_{i}u_{s},\phi)|+ |(\gb^{i}_{s} u_{s},D_{i}\phi)|+ |(c_{s}u_{s},\phi)|)\,ds $$ \begin{equation} \label{8.12.3} \leq n\|u\|_{\cL_{p}((0,\tau),W^{1}_{p})} \|\phi\|_{W^{1}_{p'}}, \end{equation} so that requirement (i) in Definition \ref{definition 3.20.01} can be dropped. \end{lemma} Proof. By having in mind partitions of unity we convince ourselves that it suffices to prove \eqref{8.12.3} under the assumption that $\phi$ has support in a ball $B_{\rho_{1}}(x_{0})$. Observe that by \eqref{1.13.1} and H\"older's inequality \begin{equation} \label{1.15.4} |(\gb^{i}_{s} u_{s},D_{i}\phi)|+|(c_{s}u_{s},\phi)|\leq N(1+|\bar{\gb}_{s}(x_{0})|+ |\bar{c}_{s}(x_{0})|)\|u_{s}\|_{W^{1}_{p}}\|\phi\|_{W^{1}_{p'}}. \end{equation} It follows again by H\"older's inequality that $$ \int_{0}^{t\wedge\tau}( |(\gb^{i}_{s} u_{s},D_{i}\phi)|+|(c_{s}u_{s},\phi)|)\,ds \leq N \chi_{t}\|u\|_{\cL_{p}((0,\tau),W^{1}_{p})} \|\phi\|_{W^{1}_{p'}}, $$ where $$ \chi_{t}= t^{1/p'} +\big(\int_{0}^{t}|\bar{\gb} _{s}(x_{0})|^{p'}ds\big)^{1/p'} +\big(\int_{0}^{t}|\bar{c} _{s}(x_{0})|^{p'}\, ds\big)^{1/p'} . $$ After that, in what concerns $\gb$ and $c$, it only remains to recall Assumption \ref{assumption 2.7.2} (iii). Similarly the integral of $ |(b^{i}_{s}D_{i}u_{s},\phi)|$ is estimated by using \eqref{1.12.2} and the lemma is proved. \begin{remark} \label{remark 9.1.1} Estimates \eqref{1.12.2} and \eqref{1.13.1} show that for any $u\in\bW^{1}_{p}$ for almost all $(\omega,s)$ the functions $b^{i}_{s}D_{i}u_{s}$, $D_{i}(\gb^{i}_{s}u_{s})$, and $c_{s}u_{s}$ are distributions on $\bR^{d}$. \end{remark} Since bounded linear operators are continuous we obtain the following. \begin{corollary} \label{corollary 3.23.1} Let $R,\tau_{n},\phi$ be as in Lemma \ref{lemma 6.27.1}. Then the operators $$ u_{t}\to \int_{0}^{t\wedge\tau_{n}} (b^{i}_{s}D_{i}u_{s},\phi)\,ds,\quad u_{t}\to \int_{0}^{t\wedge\tau_{n}} (\gb^{i}_{s}u_{s},D_{i}\phi)\,ds, $$ $$ u_{t}\to \int_{0}^{t\wedge\tau_{n}} (c_{s}u_{s}, \phi)\,ds $$ are continuous as operators from $\bW^{1}_{p}(\tau)$ to $\cL_{p}(\opar0,\tau_{n}\cbrk)$ for any $n$. \end{corollary} This result will be used in Section \ref{section 6.9.5}. Now we prove Theorem \ref{theorem 3.11.1} in a particular case. \begin{lemma} \label{lemma 3.11.1} Let $\gb^{i}$, $b^{i}$, and $c$ be independent of $x$ and let $u_{0}=0$. Then the assertion of Theorem \ref{theorem 3.11.1} holds, naturally, with $\lambda_{0}=\lambda_{0}(d,\delta, p, \rho_{0},\kappa )$ (independent of $\rho_{1}$). \end{lemma} Proof. First let $c\equiv0$. We want to use the It\^o-Wentzell formula to get rid of the first-order terms. Observe that \eqref{2.6.4} reads as $$ du_{t}=(\Lambda_{t}^{k}u_{t}+g^{k}_{t})\,dw^{k}_{t} $$ \begin{equation} \label{6.28.1} +\big[D_{i}(a^{ij}_{t}D_{j}u_{t}+(\gb^{i}_{t} + b^{i}_{t})u_{t}+f^{i}_{t})+f^{0}_{t}- \lambda u_{t}\big]\,dt, \quad t\leq \tau. \end{equation} Recall that from the start (see Definition \ref{definition 3.20.01}) it is assumed that $u \in \bW^{ 1}_{p}(\tau)$. Then one can find a predictable set $A\subset\opar0,\tau\cbrk$ of full measure such that $I_{A}f^{j}$, $j=0,1,...,d$, $I_{A} g$, and $I_{A}D_{i}u$, $i=1,...,d$, are well defined as $\cL_{p}$-valued predictable functions satisfying $$ \int_{0}^{\tau}I_{A}\big(\sum_{j=0}^{d}\|f^{j}_{t}\|^{p}_{\cL_{p}} + \|g_{t}\|^{p}_{\cL_{p}}+ \|Du_{t}\|^{p}_{\cL_{p}}\big)\,dt<\infty. $$ Replacing $f^{j}$, $g$, and $D_{i}u$ in \eqref{6.28.1} with $I_{A}f^{j}$, $I_{A}g $, and $I_{A}D_{i}u$, respectively, will not affect \eqref{6.28.1}. Similarly one can treat the term $h_{t}= (\gb^{i}_{t}+b^{i}_{t})u_{t}$ for which $$ \int_{0}^{T\wedge\tau}\|h_{t}\|_{\cL_{p}}\,dt<\infty $$ (a.s.) for each $T\in\bR_{+}$, owing to Assumption \ref{assumption 2.7.2} and the fact that $u\in\bL_{p}( \tau)$. After these replacements all terms on the right in \eqref{6.28.1} will be of class $\frD^{1} $ and $\frD^{2} $ as appropriate since $a$ and $\sigma$ are bounded (see the definition of $\frD^{1} $ and $\frD^{2} $ in \cite{Kr09_4}). This allows us to apply Theorem 1.1 of \cite{Kr09_4} and for $$ B_{t}^{i}=\int_{0}^{t} (\gb^{i}_{s}+b^{i}_{s})\,ds,\quad \hat{u}_{t}(x)=u_{t}(x-B_{t}) $$ obtain that \begin{equation} \label{6.28.8} d\hat{u}_{t}=\big[D_{i}(\hat{a}^{ij}_{t}D_{j}\hat{u}_{t} ) - \lambda \hat{u}_{t}+D_{i}\hat{f}^{i}_{t}+\hat{f}^{0}_{t} \big]\,dt+(\hat{\Lambda}^{k}_{t}\hat{u}_{t}+\hat{g}^{k}_{t}) \,dw^{k}_{t}, \end{equation} where $\hat{\Lambda}^{k}_{t}=\hat{\sigma}^{ik}_{t}D_{i}+ \hat{\nu}^{k}_{t}$ and $$ (\hat{a}^{ij}_{t},\hat{\sigma}^{ik}_{t}, \hat{\nu}^{k}_{t} \hat{f}^{j}_{t},\hat{g}^{k}_{t})(x) =(a^{ij}_{t}, \sigma^{ik}_{t},\nu^{k}_{t}, f^{j}_{t}, g^{k}_{t})(x-B_{t}). $$ Obviously, $\hat{u}$ is in $\bW^{1}_{p}(\tau)$ and its norm coincides with that of $u$. Equation \eqref{6.28.8} shows that $\hat{u}\in\cW^{1}_{p}(\tau)$. Next observe that owing to \eqref{2.9.5}, for any $\omega\in\Omega,\rho\in(0,\rho_{0}], t\geq0$, and $i,j=1,...,d$ we have $$ \rho^{-2d-2}\int_{t}^{t+\rho^{2}}\bigg( \sup_{x\in \bR^{d}}\int_{B_{\rho}(x)}\int_{B_{\rho}(x)} |\hat{a}^{ij}_{s}( y)-\hat{a}^{ij}_{s}( z)|\,dydz\bigg) \,ds\leq\gamma_{a}, $$ which in terms of \cite{Kr09} implies that the couple $(\hat{a},\hat{\sigma})$ is $(\varepsilon, \varepsilon )$-regular at any point of $\bR_{+}\times\bR^{d}$ for any $\varepsilon\in(0,\rho_{0}]$. Then owing to our Assumptions \ref{assumption 2.7.2} (ii) and \ref{assumption 2.4.1} one can choose $\varepsilon=\varepsilon(\delta,\kappa) \in(0,\rho_{0}]$ so that Assumption 2.2 of \cite{Kr09} is satisfied. By Theorem 2.2 of \cite{Kr09} if Assumption \ref{assumption 2.4.01} is satisfied with $\gamma_{a}=\gamma _{a}(d,\delta,p)>0$, specified in its proof, and if $\lambda\geq\lambda_{0}(d,\delta,p,\kappa, \rho_{0})\geq1$, then $$ \lambda\|\hat u\|^{2}_{\bL _{p}(\tau)}+\|D\hat u\|^{2}_{\bL _{p}(\tau)} \leq N\big(\sum_{i=1}^{d}\|\hat f^{i}\|^{2}_{\bL _{p}(\tau)} +\|\hat g\|^{2}_{\bL _{p}(\tau)} + \lambda^{-1}\|\hat f^{0}\|^{2}_{\bL _{p}(\tau)}\big) , $$ where $N=N(d,\delta,p,\kappa, \rho_{0})$. This coincides with \eqref{3.11.2} and proves the lemma in case $c\equiv0$. In the general case observe that owing to Assumption \ref{assumption 2.7.2} (iii) there exists a sequence of stopping times $\tau_{n}\uparrow\tau$ such that $$ \int_{0}^{\tau_{n}}c_{s}\,ds\leq n. $$ Clearly, if we can prove \eqref{3.11.2} with $\tau_{n}$ in place of $\tau$, then by passing to the limit we will get \eqref{3.11.2} as is. Therefore, without losing generality we assume that $$ \sup_{\Omega}\int_{0}^{\infty}c_{s}\,ds<\infty. $$ Then introduce $$ \xi_{t}=\exp(\int_{0}^{t}c_{s}\,ds). $$ By the above argument we have $\bar{u}:= \xi u\in \bW^{1}_{p}(\tau)$ and $$ d\bar{u}_{t}= \big[D_{i}(a^{ij}_{t}D_{j}\bar{u}_{t}+[\gb^{i}_{t} + b^{i}_{t}]\bar{u}_{t}+\xi_{t}f^{i}_{t})+\xi_{t}f^{0}_{t}- \lambda \bar{u}_{t} \big]\,dt+(\Lambda^{k}_{t}\bar{u}_{t} +\xi_{t}g^{k}_{t})\,dw^{k}_{t}, \quad t\leq \tau. $$ By the above result for any stopping time $\tau'\leq \tau$ $$ \lambda^{p/2}\|\xi u\|^{p}_{\bL _{p}(\tau')}+ \|\xi Du\|^{p}_{\bL _{p}(\tau')}= \lambda^{p/2}\|\bar{u}\|^{p}_{\bL _{p}(\tau')}+ \|D\bar{u}\|^{p}_{\bL _{p}(\tau')} $$ \begin{equation} \label{3.11.02} \leq N\big(\sum_{i=1}^{d}\|\xi f^{i}\|^{p}_{\bL _{p}(\tau')} +\|\xi g\|^{p}_{\bL _{p}(\tau')} + \lambda^{-p/2}\|\xi f^{0}\|^{p}_{\bL _{p}(\tau')}\big). \end{equation} If needed, one can enlarge the original probability space in such a way that there will exist an exponentially distributed, with parameter one, random variable $\eta$ independent of $\{\cF_{t},t\geq0\}$. We assume that the enlargement is not needed and define $$ \phi_{t}=p\int_{0}^{t}c_{s}\,ds,\quad \psi_{s}=\tau\wedge\inf\{t\geq 0: \phi_{t}\geq s\}, \quad\tau' =\psi_{\eta}. $$ Notice that $$ \{\omega:\psi_{s}> t\}=\{\omega:\tau>t, \phi_{t}< s\}. $$ Hence $$ \{\omega:\tau'> t\}=\{\omega:\tau>t, \phi_{t}< \eta\}. $$ It follows that $\tau'$ is a stopping time with respect to $\cF_{t}\vee\sigma(\eta)$. Furthermore, for any nonnegative predictable (relative to the original filtration $\cF_{t}$) process $h_{t}$ we have $$ E\int_{0}^{\tau'}h_{t}\,dt=\int_{0}^{\infty} Eh_{t}E\{I_{\tau'>t}\mid \cF_{t}\}\,dt $$ $$ =\int_{0}^{\infty} Eh_{t} I_{\tau >t}e^{- \phi_{t}}\,dt=E\int_{0}^{\tau}h_{t} \xi^{-p}_{t}\,dt. $$ This and \eqref{3.11.02} immediately lead to \eqref{3.11.2} and the lemma is proved. To proceed further take $\bar{\gb}, \bar{b}$, and $\bar{c}$ from \eqref{6.28.3}. From Lemma 4.2 of \cite{Kr09_1} and Assumption \ref{assumption 3.11.1} it follows that, for $h_{ t}=\bar{\gb}_{ t},\bar{b}_{ t}, \bar{c}_{ t}$, it holds that $|D^{n}h_{ t}|\leq M_{n} $, where $M_{n}=M_{n} (n, d,\rho_{1},K)\geq1$ and $ D^{n}h_{ t }$ is any derivative of $h_{ t}$ of order $n\geq1$ with respect to $x$. By Corollary 4.3 of \cite{Kr09_1} we have $|h_{ t}( x )|\leq K(t)(1+|x|)$, where for each $\omega$ the function $K(t)=K(\omega,t)$ is locally integrable with respect to $t$ on $\bR_{+}$. Owing to these properties the equation \begin{equation} \label{2.8.1} x_{t}=x_{0}-\int_{t_{0}}^{t}(\bar{\gb}_{ s}+\bar{b}_{ s}) ( x_{s})\,ds,\quad t \geq t_{0} , \end{equation} for any ($\omega$ and) $ (t_{0},x_{0}) \in \bR^{d+1 }_{+} $ has a unique solution $x_{t}=x_{t_{0},x_{0},t} $. Obviously, the process $x_{t_{0},x_{0},t}$, $t\geq t_{0}$, is $\cF_{t}$-adapted. Next, for $i=1,2$ set $\chi^{(i)}(x)$ to be the indicator function of $B_{\rho_{1}/i}$ and introduce $$ \chi^{(i)}_{t_{0},x_{0},t}(x)=\chi^{(i)}(x-x_{t_{0},x_{0},t}) I_{t\geq t_{0}}. $$ By using the above results and reproducing the proofs of Lemma 5.5 of \cite{Kr10_1}, where $p=2$ and SPDEs are treated, and Lemma 5.8 of~\cite{Kr10_2}, where $p$ is general but only PDEs are considered, we easily obtain the following. \begin{lemma} \label{lemma 3.14.1} Suppose that Assumption \ref{assumption 2.4.01} is satisfied with $\gamma_{a}=\gamma_{a}(d,\delta,p)$ taken from Lemma \ref{lemma 3.11.1}. Assume that we are given a function $u $ which is a solution of \eqref{2.6.4} with some $f^{j},g \in\bL_{p}(\tau)$, and $\lambda\geq\lambda_{0}= \lambda_{0}(d,\delta,p, \rho_{0},\kappa )$, where $\lambda_{0}(d,\delta, p, \rho_{0},\kappa )$ is taken from Lemma~\ref{lemma 3.11.1}. Take $ (t_{0},x_{0}) \in \bR^{d+1}_{+}$ and assume that $u_{t}=0$ if $t\leq t_{0}\wedge \tau$. Then $$ \lambda^{p/2} \|\chi^{(2)}_{t_{0},x_{0}}u \|^{p}_{\bL _{p}(\tau)}+ \|\chi^{(2)}_{t_{0},x_{0}}Du\|^{p}_{\bL _{p}(\tau)} $$ $$ \leq N\big(\sum_{i=1}^{d}\|\chi^{(1)}_{t_{0},x_{0}} f^{i}\|^{p}_{\bL _{p}(\tau)}+\| \chi^{(1)}_{t_{0},x_{0}}g\|^{p}_{\bL_{p}(\tau)}\big) +N\lambda^{-p/2}\|\chi^{(1)}_{t_{0},x_{0}}f^{0}\|^{p}_{\bL _{p}(\tau)} $$ $$ +N\gamma_{b}^{p/q} \| \chi^{(1)}_{t_{0},x_{0}} Du \|_{\bL_{p}(\tau)}^{p}+ N^{*} \lambda^{-p/2}\| \chi^{(1)}_{t_{0},x_{0}} Du \|_{\bL_{p}(\tau)}^{p} $$ \begin{equation} \label{3.14.2} + N^{*} \| \chi^{(1)}_{t_{0},x_{0}} u \|_{\bL_{p}(\tau)}^{p} +N^{*}\lambda^{-p/2}\sum_{i=1}^{d}\|\chi^{(1)}_{t_{0},x_{0}} f^{i}\|^{p}_{\bL _{p}(\tau)}, \end{equation} where $N$ is a constant depending only on $d,\delta$, $p$, $\rho_{0}$, and $\kappa$ and $N^{*}$ depends only on the same objects, $\gamma_{b}$, $\rho_{1}$, and $K$. \end{lemma} Upon integrating through equation \eqref{3.14.2} with respect to $x_{0}$ and repeating the arguments in the proofs of Lemma 5.6 of \cite{Kr10_1} or Lemma 5.9 of \cite{Kr10_2} we obtain the following result in which $M_{1}( d,\rho_{1},K)$ is the constant introduced before Lemma \ref{lemma 3.14.1}. \begin{lemma} \label{lemma 3.14.3} Suppose that Assumption \ref{assumption 2.4.01} is satisfied with $\gamma_{a}=\gamma_{a}(d,\delta,p) $ taken from Lemma \ref{lemma 3.11.1}. Assume that we are given a function $u $ which is a solution of \eqref{2.6.4} with some $f^{j},g \in\bL_{p}(\tau)$, and $\lambda\geq\lambda_{0}= \lambda_{0}(d,\delta,p, \rho_{0},\kappa )$, where $\lambda_{0}(d,\delta, p, \rho_{0},\kappa )$ is taken from Lemma~\ref{lemma 3.11.1}. Take an $s_{0}\in\bR_{+}$ and assume that $u_{t}=0$ if $t\leq s_{0}\wedge\tau$. Then for $I_{s_{0}}:= I_{(s_{0},t_{0})}$, where $t_{0}= s_{0}+M_{1}^{-1} $, we have $$ \lambda^{p/2}\| I_{s_{0}}u\|^{p}_{\bL _{p}(\tau) }+ \|I_{s_{0}} Du\|^{p}_{\bL _{p}(\tau) } \leq N\big( \sum_{i=1}^{d}\| I_{s_{0}}f^{i}\|^{p}_{\bL _{p}(\tau) }+\| I_{s_{0}}g\|^{p}_{\bL _{p}(\tau) }\big) $$ $$ +N\lambda^{-p/2 }\|I_{s_{0}} f^{0}\|^{p}_{\bL _{p} (\tau)} +N\gamma_{b} ^{p/q} \| I_{s_{0}} Du\|_{\bL_{p}(\tau) }^{p} $$ \begin{equation} \label{3.14.5} + N^{*} \lambda^{-p/2}\| I_{s_{0}} Du\|_{\bL_{p}(\tau) }^{p} + N^{*} \| I_{s_{0}} u\|_{\bL_{p}(\tau) }^{p} +N^{*}\lambda^{-p/2}\sum_{i=1}^{d}\| I_{s_{0}}f^{i}\|^{p}_{\bL _{p}(\tau) }, \end{equation} where $N$ is a constant depending only on $d,\delta$, $p$, $\rho_{0}$, and $\kappa$ and $N^{*}$ depends only on the same objects, $\gamma_{b}$, $\rho_{1}$, and $K$. \end{lemma} {\bf Proof of Theorem \ref{theorem 3.11.1}}. First we show how to choose an appropriate $\gamma_{b}=\gamma_{b}(d,\delta,p,\rho_{0},\kappa)$. Call $N_{0}$ the constant factor of $\gamma_{b}^{p/q} \| I_{s_{0}} Du\|_{\bL_{p}(\tau) }^{p}$ in \eqref{3.14.5} and choose a $\gamma_{b}\in(0,1]$ in such a way that $N_{0} \gamma^{p/q}_{b}\leq 1/2$. Then under the assumptions of Lemma \ref{lemma 3.14.3} we have $$ \lambda^{p/2}\| I_{s_{0}}u\|^{p}_{\bL _{p}(\tau) }+ \|I_{s_{0}} Du\|^{p}_{\bL _{p}(\tau) } \leq N \big(\sum_{i=1}^{d}\| I_{s_{0}}f^{i}\|^{p}_{\bL _{p}(\tau) }+\| I_{s_{0}}g\|^{p}_{\bL _{p}(\tau) }\big) $$ $$ +N\lambda^{-p/2 }\|I_{s_{0}} f^{0}\|^{p}_{\bL _{p} (\tau)} + N^{*} \lambda^{-p/2}\| I_{s_{0}} Du\|_{\bL_{p}(\tau) }^{p} $$ \begin{equation} \label{8.28.1} + N^{*} \| I_{s_{0}} u\|_{\bL_{p}(\tau) }^{p} +N^{*}\lambda^{-p/2}\sum_{i=1}^{d}\| I_{s_{0}}f^{i}\|^{p}_{\bL _{p}(\tau) }. \end{equation} To proceed further assume that \begin{equation} \label{8.28.3} u_{0}=0. \end{equation} After $\gamma_{b}$ has been fixed we recall that $M_{1}=M_{1} ( d,\rho_{1},K)$ and we take a $\zeta\in C^{\infty}_{0}(\bR)$ with support in $(0,M_{1}^{-1})$ such that \begin{equation} \label{3.15.1} \int_{-\infty}^{\infty}\zeta^{p}(t)\,dt=1. \end{equation} For $s\in\bR $ define $\zeta^{s}_{t}=\zeta(t-s)$, $u^{s}_{t}( x)=u_{t}(x)\zeta^{s}_{t}$. Obviously, $u^{s}_{t}=0$ if $0\leq t\leq s_{+}\wedge\tau$. Therefore, we can apply \eqref{8.28.1} to $u^{s}_{t}$ by taking $s_{0}=s_{+}$ and observing that $$ du^{s}_{t}=(L_{t}u^{s}_{t}-\lambda u^{s}_{t} +D_{i}(\zeta^{s}_{t}f^{i}_{t})+\zeta^{s}_{t}f^{0}_{t}+u_{t}(\zeta^{s}_{t})')\,dt +(\Lambda^{k}_{t}u^{s}_{t}+\zeta^{s}_{t}g^{k}_{t})\,dw^{k}_{t}, \quad t\leq\tau. $$ We also use the fact that for $t\geq0$, as is easy to see, $I_{s_{+}}(t)\zeta^{s}_{t}=\zeta^{s}_{t}$. Then for and $\lambda\geq\lambda_{0}= \lambda_{0}(d,\delta,p, \rho_{0},\kappa )$, where $\lambda_{0}(d,\delta, p, \rho_{0},\kappa )$ is taken from Lemma~\ref{lemma 3.11.1}, we obtain $$ \lambda^{p/2}\| \zeta^{s} u\|^{p}_{\bL _{p}(\tau) }+ \|\zeta^{s} Du\|^{p}_{\bL _{p}(\tau) } \leq N \big(\sum_{i=1}^{d}\| \zeta^{s} f^{i}\|^{p}_{\bL _{p}(\tau) } +\| \zeta^{s}g\|^{p}_{\bL _{p}(\tau) }\big) $$ $$ +N\lambda^{-p/2 }(\|\zeta^{s} f^{0}\|^{p}_{\bL _{p} (\tau)} +\|(\zeta^{s})' u\|^{p}_{\bL _{p} (\tau)}) $$ \begin{equation} \label{8.28.2} + N^{*} \lambda^{-p/2}\| \zeta^{s} Du\|_{\bL_{p}(\tau) }^{p} + N^{*} \| \zeta^{s} u\|_{\bL_{p}(\tau) }^{p} +N^{*}\lambda^{-p/2}\sum_{i=1}^{d}\| \zeta^{s} f^{i}\|^{p}_{\bL _{p}(\tau) }. \end{equation} We integrate through this relation with respect to $s\in\bR $, use \eqref{3.15.1} and $$ \int_{-\infty}^{\infty}|(\zeta^{s}_{t})'|^{p}\,ds= \int_{-\infty}^{\infty}| \zeta'(t) |^{p}\,dt=N^{*}. $$ Then we conclude $$ \lambda^{p/2}\| u\|^{p}_{\bL _{p}(\tau) }+ \| Du\|^{p}_{\bL _{p}(\tau) } \leq N_{1} \big(\sum_{i=1}^{d}\| f^{i}\|^{p}_{\bL _{p}(\tau) } +\| g\|^{p}_{\bL _{p}(\tau) }\big) $$ $$ + N_{1}\lambda^{-p/2 } \| f^{0}\|^{p}_{\bL _{p} (\tau)} + N^{*} _{1} \lambda^{-p/2}\| Du\|_{\bL_{p}(\tau) }^{p} $$ $$ + N^{*} _{1} \| u\|_{\bL_{p}(\tau) }^{p} +N^{*} _{1}\lambda^{-p/2}\sum_{i=1}^{d}\| f^{i}\|^{p}_{\bL _{p}(\tau) }. $$ Without losing generality we assume that $N_{1}\geq1$ and we show how to choose $\lambda_{0}=\lambda_{0}( d,\delta,p,\rho_{0},\rho_{1},\kappa,K)\geq1$. Above we assumed that $\lambda\geq\lambda_{0}(d,\delta,p,\rho_{0} ,\kappa)$, where $\lambda_{0}(d,\delta,p,\rho_{0},\kappa)$ is taken from Lemma \ref{lemma 3.11.1}. Therefore, we take $$ \lambda_{0}=\lambda_{0}( d,\delta,p,\rho_{0},\rho_{1},\kappa,K)\geq \lambda_{0}(d,\delta,p,\rho_{0},\kappa) $$ such that $\lambda_{0}^{p/2}\geq 2N^{*}_{1} $ (recall that $N_{1}^{*}=N_{1}^{*} (d,\delta,p,\rho_{0},\rho_{1},\kappa,K)$). Then we obviously come to \eqref{3.11.2} (with $u_{0}=0$). A standard method to remove assumption \eqref{8.28.3} by subtracting from $u$ the solution of the heat equation $dv_{t}=(\Delta v_{t}-v_{t})\,dt$ with initial data $u_{0}$ does not work because it leads to subtracting the terms $D_{i}(\gb^{i}v)+b^{i}D_{i}v$, which one should include into the free terms $D_{i}f^{i}+f^{0}$ in the equation. Generally, this is impossible because we only know that $D_{i}v\in\bL_{p}(\tau)$ and if we multiply $D_{i}v$ by an arbitrary function of $x$ with linear growth, the inclusion may fail. Therefore, we use a different method. The idea is to shift all data along the time axis by 1, consider our equations on $\opar1,\hat \tau \cbrk$, where $\hat \tau =1+\tau$, and supplement this equation with an equation for $t\in[0,1]$ with zero initial data and such that the value of its solution at time 1 would coincide with $u_{0}$. Then the two equations combined would give an equation on $\opar0,\hat \tau \cbrk$ with zero initial condition, which would allow us to apply the above result. Formally, we need to have Wiener processes on $[0,\infty)$ and after shifting they will be defined only on $[1,\infty)$ (and satisfy $w^{k}_{1}=0$). Therefore, we augment if needed our probability space in such a way that we may assume that there are Wiener processes $\bar{w}^{1}_{t}, \bar{w}^{2}_{t},...$, $t\geq0$, independent of $\{\cF_{s},s\geq0\}$. Then define $\cF^{\bar w}_{t}$ as the completion of $\sigma(\bar{w}_{s}:s\leq t)$, $$ \hat\cF _{t}=\cF_{0}\vee\cF^{\bar w}_{t},\quad t\in[0,1], \quad \hat\cF _{t}=\cF_{t-1}\vee\cF^{\bar w}_{1},\quad t\geq1, $$ $$ \hat w^{ k}_{t}=\bar{w}^{k}_{t},\quad t\in[0,1],\quad \hat w^{ k}_{t}=\bar{w}^{k}_{1}+w^{k}_{t-1}\quad t\geq1, \quad \hat\tau=1+\tau, $$ and for $t\geq1$ define the coefficients and the free terms by following the example $\hat{a}^{ij}_{t}=a^{ij}_{t-1}$. Next, take the function $v$ from Definition \ref{definition 8.10.1} and for $t\in[0,1]$ set $$ \hat{a}^{ij}_{t}=\delta^{ij},\quad \hat f^{i}_{t}=-2tD_{i}v_{1-t}, \quad\hat f^{0}_{t}=(1+t+\lambda t)v_{1-t}, $$ where $\lambda\geq\lambda_{0}$ with $\lambda_{0}$ determined in the first part of the proof. We define all other coefficients with hats and the free terms $\hat{g}^{k}_{t}$ to be zero for $t\in[0,1]$. Notice that for $\hat{u}_{t}=tv_{1-t}$, $t\in[0,1]$, we have $$ d\hat{u}_{t}=\big[D_{i}(\hat{a}^{ij}_{t}D_{j}\hat{u}_{t}+ \hat{f}^{i}_{t})+\hat{f}^{0}_{t}-\lambda \hat{u}_{t}\big]\,dt. $$ Moreover, $\hat{u}_{0}=0$, $\hat{u}_{1}=u_{0}$, and $\hat{u}_{t}$ is $\hat{\cF}_{t}$-adapted. Therefore, naturally we define $\hat{u}_{t}=u_{t-1}$ for $t\geq 1$. It is easy to see that if we construct the operators $\hat{L}_{t}$ and $\hat{\Lambda}^{k}_{t}$ from the coefficients with hats, then $$ d\hat u_{t}=(\hat L_{t}\hat u_{t}-\lambda \hat u_{t}+D_{i}\hat f^{i}_{t}+\hat f^{0}_{t})\,dt +(\hat \Lambda^{k}_{t}\hat u_{t}+\hat g^{k}_{t})\,d\hat w^{k}_{t}, \quad t\leq\hat \tau. $$ By the first part of the proof $$ \lambda\|u\|^{2}_{\bL _{p}(\tau)}+\|Du\|^{2}_{\bL _{p}(\tau)} \leq\lambda\|\hat u\|^{2}_{\bL _{p}(\hat\tau)}+ \|D\hat u\|^{2}_{\bL _{p}(\hat\tau)} $$ $$ \leq N\big(\sum_{i=1}^{d}\|\hat f^{i}\|^{2}_{\bL _{p}(\hat\tau)} +\|\hat g\|^{2}_{\bL _{p}(\hat\tau)} \big) +N\lambda^{-1}\|\hat f^{0}\|^{2}_{\bL _{p}(\hat \tau)} $$ $$ \leq N\big(\sum_{i=1}^{d}\| f^{i}\|^{2}_{\bL _{p}( \tau)} +\| g\|^{2}_{\bL _{p}(\ \tau)} \big) +N\lambda^{-1}\| f^{0}\|^{2}_{\bL _{p}(\tau)} $$ $$ +N(\|v\|^{2}_{\bL_{p}}+\|Dv\|^{2}_{\bL_{p}}). $$ It only remains to notice that the last term is dominated by $N\|u_{0}\|_{\text{\rm tr}\, \cW^{1}_{p}}^{2}$. The theorem is proved. \mysection{Proof of Theorem \protect\ref{theorem 3.16.1}} \label{section 6.9.5} Throughout this section we suppose that the assumptions of Theorem \ref{theorem 3.16.1} are satisfied. Owing to Theorem \ref{theorem 3.11.1}, implying that the solution in $\bW^{1}_{p}(\tau)$ is unique, and having in mind setting all data equal to zero for $t>\tau$, we see that without loss of generality we may assume that $\tau=\infty$. Set $$ \bL_{p}=\bL_{p}( \infty),\quad \bW^{1}_{p}=\bW^{1}_{p}( \infty). $$ We need two auxiliary results. \begin{lemma} \label{lemma 9.2.1} For any $T,R\in(0,\infty)$ (and $\omega$), we have \begin{equation} \label{9.2.2} \int_{0}^{T}\int_{B_{R}}(|\gb_{s}(x)|^{p'} +|b_{s}(x)|^{p'} +c_{s}^{p'}(x)) \,dxds<\infty. \end{equation} \end{lemma} This lemma is proved in the same way as Lemma 6.1 of \cite{Kr10_2} on the basis of Assumptions \ref{assumption 2.7.2} (iii) and \ref{assumption 3.11.1} and the fact that $q\geq p'$. The solution of our equation will be obtained as the weak limit of the solutions of equations with cut-off coefficients. Therefore, the following result is relevant. \begin{lemma} \label{lemma 3.23.2} Let $\phi\in C^{\infty}_{0}$, $u^{m}$, $u\in \bW^{1}_{p}$, $m=1,2,...$, be such that $u^{m}\to u$ weakly in $\bW^{1}_{p}$. For $m=1,2,...$ define $\chi_{m}(t)=(-m)\vee t\wedge m$, $\gb^{i}_{mt}=\chi_{m}(\gb^{i}_{t})$, $b^{i}_{mt}=\chi_{m}(b^{i}_{t})$, and $c_{mt}=\chi_{m}(c_{t})$. Then there is a sequence of bounded stopping times $\tau_{n} \to\infty$ such that, for any $n$, the functions \begin{equation} \label{4.19.6} \int_{0}^{t} (b^{i}_{ms}D_{i}u^{m}_{s},\phi)\,ds,\quad \int_{0}^{t} (\gb^{i}_{ms}u^{m}_{s},D_{i}\phi) \,ds,\quad \int_{0}^{t} (c_{ms}u^{m}_{s}, \phi) \,ds \end{equation} converge weakly in the space $\cL_{p}(\opar0,\tau_{n}\cbrk)$ as $m\to\infty$ to \begin{equation} \label{8.12.1} \int_{0}^{t} (b^{i}_{s}D_{i}u_{s},\phi)\,ds, \quad \int_{0}^{t} (\gb^{i}_{s}u_{s},D_{i}\phi)\,ds,\quad \int_{0}^{t} (c_{s} u_{s},\phi)\,ds, \end{equation} respectively. \end{lemma} Proof. Let $R$ be such that $\phi(x)=0$ for $|x|\geq R$. We take $\tau_{n}\to\infty$ such that each of them is bounded, they are smaller than the ones from Lemma \ref{lemma 6.27.1}, and are such that the left hand side of \eqref{9.2.2} with $T=\tau_{n}$ is less than $n$. By Corollary \ref{corollary 3.23.1} and by the fact that (strongly) continuous operators are weakly continuous we obtain that $$ \int_{0}^{t} (b^{i}_{s}D_{i}u^{m}_{s},\phi) \,ds \to \int_{0}^{t} (b^{i}_{s}D_{i}u_{s},\phi) \,ds $$ as $m\to\infty$ weakly in the space $\cL_{p}(\opar0,\tau_{n} \cbrk)$ for any $n$. Therefore, in what concerns the first function in \eqref{4.19.6}, it suffices to show that $$ \int_{0}^{t} (D_{i}u^{m}_{s},(b^{i}_{s}-b^{i}_{ms})\phi) \,ds \to0 $$ weakly in $\cL_{p}(\opar0,\tau_{n} \cbrk)$. In other words, it suffices to show that for any $\xi\in \cL_{p'}(\opar0,\tau_{n} \cbrk)$ $$ E\int_{0}^{\tau_{n}}\xi_{t} \big(\int_{0}^{t} (D_{i}u^{m}_{s},(b^{i}_{s}-b^{i}_{ms})\phi) \,ds \big)\,dt\to0. $$ This relation is rewritten as \begin{equation} \label{4.21.1} E \int_{0}^{\tau_{n}} (D_{i}u^{m}_{s},\eta_{s}(b^{i}_{s}-b^{i}_{ms})\phi) \,ds\to0, \end{equation} where $$ \eta_{s}:=\int_{s}^{\tau_{n} }\xi_{t}\,dt. $$ Observe that by the choice of $\tau_{n}$ we have $$ E\int_{0}^{\tau_{n}}|\eta_{s}|^{p'} \int_{|x|\leq R} |b_{s}(x)|^{p'}\,dxds \leq E\sup_{t\leq\tau_{n}}|\eta_{s}|^{p'} \int_{0}^{\tau_{n}} \int_{|x|\leq R} |b_{s}(x)|^{p'}\,dxds $$ $$ \leq nE\big(\int_{0}^{\tau_{n}}|\xi_{s}|\,ds\big)^{p'}<\infty. $$ It follows by the dominated convergence that $ \eta_{s}(b^{i}_{s}-b^{i}_{ms})\phi\to0 $ as $m\to\infty$ strongly in $\bL_{p'}( \tau_{n} )$. By assumption $Du^{m}\to Du$ weakly in $\bL_{p}( \tau_{n} )$. This implies \eqref{4.21.1}. Similarly, one proves our assertion about the remaining functions in \eqref{4.19.6}. The lemma is proved. {\bf Proof of Theorem \ref{theorem 3.16.1}}. Recall that we may assume that $ \tau=\infty$. Since the case $p=2$ is dealt with in \cite{Kr10_1} (under much milder assumptions), we also assume that $p>2$. Define $\gb_{mt}$, $b_{mt}$, and $c_{mt}$ as in Lemma \ref{lemma 3.23.2} and consider equation \eqref{2.6.4} with $\gb_{mt}$, $b_{mt}$, and $c_{mt}$ in place of $\gb_{t}$, $b_{t}$, and $c_{t}$, respectively. Obviously, $\gb_{mt}$, $b_{mt}$, and $c_{mt}$ satisfy Assumption \ref{assumption 3.11.1} with the same $\gamma_{b}$ and $K$ as $\gb_{t}$, $b_{t}$, and $c_{t}$ do. By Theorem \ref{theorem 3.11.1} and the method of continuity for $\lambda\geq\lambda_{0}(d,\delta,p,\kappa, \rho_{0},\rho_{1},K)$ there exists a unique solution $u^{m} $ of the modified equation on $\bR$. By Theorem \ref{theorem 3.11.1} we also have $$ \|u^{m}\|_{\bL_{p} }+\|Du^{m}\|_{\bL_{p} }\leq N, $$ where $N$ is independent of $m$. Hence the sequence of functions $u^{m} $ is bounded in the space $\bW^{1}_{p}$ and consequently has a weak limit point $u\in \bW^{1}_{p}$. For simplicity of presentation we assume that the whole sequence $u^{m} $ converges weakly to $u$. Take a $\phi\in C^{\infty}_{0}$. Then by Lemma \ref{lemma 3.23.2} for appropriate $\tau_{n}$ we have that the functions \eqref{4.19.6} converge to \eqref{8.12.1} weakly in $\cL_{p}(\opar0,\tau_{n}\cbrk)$ as $m\to\infty$ for any $n$. Owing to \eqref{9.3.3} and the fact that bounded linear operators are weakly continuous, the stochastic terms in the equations for $u^{m}_{t}$ also converge weakly in $\cL_{p}(\opar0,\tau_{n}\cbrk)$ as $m\to\infty$ for any $n$. Obviously, the same is true for $(u^{m}_{ t},\phi)\to(u_{t},\phi)$ and the remaining terms entering the equation for $ u^{m}_{t}$. Hence, by passing to the weak limit in the equation for $u^{m}_{ t}$ we see that for any $\phi\in C^{\infty}_{0}$ equation \eqref{3.16.7} holds for {\em almost any\/} $(\omega,t)$. Until this moment Assumption \ref{assumption 3.16.1} was not needed. We will need it in order to be able to apply Theorem 3.1 of \cite{KR} and find an appropriate modification of $u_{t}$. Take a $\psi\in C^{\infty}_{0}$ and observe that $u\psi\in\bW^{1}_{2}(T)$ and $g\psi\in\bL_{2}(T)$ for any $T\in(0,\infty)$ which implies that $$ m^{\psi}_{t}:=u_{0}\psi+ \sum_{k=1}^{\infty}\int_{0}^{t} \psi (\Lambda^{k}_{s}u_{s} +g^{k}_{s} )\,dw^{k}_{s} $$ is well defined as an $\cL_{2}$-valued continuous martingale such that for any $\phi\in \cL_{2}$ with probability one \begin{equation} \label{1.26.1} (m^{\psi}_{t},\phi)=(u_{0}\psi,\phi)+ \sum_{k=1}^{\infty}\int_{0}^{t}\big( \psi (\Lambda^{k}_{s}u_{s} +g^{k}_{s} ),\phi\big)\,dw^{k}_{s} \end{equation} for all $t\in\bR_{+}$. Notice that for any $\phi\in C^{\infty}_{0}$ \begin{equation} \label{9.3.4} (u_{t}\psi,\phi)=\int_{0}^{t}(u^{*}_{s}\psi ,\phi)\,ds+ (m^{\psi}_{t},\phi) \end{equation} for almost all $(\omega,t)$, where $u^{*}_{s}$ is a function with values in the space of distributions on $\bR^{d}$ defined by $$ u^{*}_{s}= L_{s}u_{s}-\lambda u_{s}+D_{i}f^{i}_{s}+f^{0}_{s} $$ (see Remark \ref{remark 9.1.1}). Next, take an $R\in(0,\infty)$ and let $W^{-1}_{p'}(B_{R})$ denote the dual space for $$ \WO^{1}_{p}(B_{R}) := W^{1}_{p}(B_{R})\cap\{v:v|_{\partial B_{R}}=0\}. $$ Estimate \eqref{1.15.4} combined with the facts that, $p'<p$ and that one can cover $B_{R}$ with finitely many balls of radius $\rho_{1}$ shows that for any $\phi\in C^{\infty}_{0}(B_{R})$ $$ |(D_{i}(\gb^{i}_{s}u_{s}),\phi)|\leq N\big(1+ \int_{B_{R+1}}|\gb_{s}|\,dx\big) \|u_{s}\|_{W^{1}_{p}} \|\phi\|_{W^{1}_{p}} , $$ where $N$ is independent of $\omega,s, u_{s},\phi$. Due to the arbitrariness of $\phi$ and the fact that $C^{\infty}_{0}(B_{R})$ is dense in $\WO^{1}_{p}(B_{R})$ we conclude that (for almost all $(\omega,s)$) we have $D_{i}(\gb^{i}_{s}u_{s})\in W^{-1}_{p'}(B_{R})$ and $$ \|D_{i}(\gb^{i}_{s}u_{s}) \|_{W^{-1}_{p'}(B_{R})} \leq N\big(1+ \int_{B_{R+1}}|\gb_{s}|\,dx\big) \|u_{s}\|_{W^{1}_{p}}. $$ Here the right-hand side is locally summable on $\bR_{+}$ to the power $p'$ (a.s.) owing to Assumption \ref{assumption 3.16.1}, H\"older's inequality, and the fact that $u\in \bW^{1}_{p}$. Similar statements are true for $b^{i}_{s}D_{i}u_{s}$, $c_{s}u_{s}$, and $u^{*}_{s}$. Now, since $u\psi\in\cL_{p}(\bR_{+},\WO^{1}_{p}(B_{R}))$ and $\WO^{1}_{p}(B_{R})$ is dense in $\cL_{2}(B_{R})$, by Theorem 3.1 of \cite{KR} we get that there exist an event $\Omega^{\psi}$ of full probability and a continuous $\cL_{2}(B_{R})$-valued $\cF_{t}$-adapted process $u^{\psi}_{t}$ such that $u^{\psi}_{t}=u_{t}\psi$ as $\cL_{2}(B_{R})$-valued functions for almost all $(\omega,t)$ and for any $\omega\in\Omega^{\psi}$, $t\in\bR_{+}$, and $\phi\in C^{\infty}_{0}(B_{R})$ we have \begin{equation} \label{9.3.7} (u^{\psi}_{t},\phi)=\int_{0}^{t}(u^{*}_{s}\psi,\phi)\,ds+ (m^{\psi}_{t},\phi). \end{equation} Take a $\psi\in C^{\infty}_{0}$ such that $\psi(x)=1$ for $|x|\leq1$ and for $k=1,2,...$ define $\psi_{k}(x)=\psi(x/k)$ and $$ \Omega'=\bigcap_{k=1}^{\infty}\Omega^{\psi_{k}}. $$ Clearly, $P(\Omega')=1$. We will further reduce $\Omega'$ in the following way. Obviously (see \eqref{1.26.1}), if $\psi',\psi''\in C^{\infty}_{0}$ and $\psi'=\psi''$ on $B_{R}$ and $\phi\in\cL_{2}$ is such that $\phi=0$ outside $B_{R}$, then with probability one we have $(m_{t}^{\psi'},\phi)= (m_{t}^{\psi''},\phi)$ for all $t$. Let $\Phi$ be the union over $n=1,2,...$ of countable subsets of $C^{\infty}_{0}(B_{n})$ each of which everywhere dense in $\cL_{2}(B_{n})$. For $\phi\in C^{\infty}_{0}$ denote $d(\phi)$ the smallest radius of the balls centered at the origin containing the support of $\phi$. Then by the above for $\phi\in C^{\infty}_{0}$ the events $$ \Omega(\phi)=\{\omega\in\Omega: (m_{t}^{\psi_{k}},\phi)= (m_{t}^{\psi_{j}},\phi),\quad\forall t\in\bR_{+}, k,j\geq d(\phi) \}, $$ $$ \Omega''=\Omega'\bigcap\bigcap_{\phi\in\Phi}\Omega(\phi) $$ have probability one. Since $m^{\psi}_{t}$ are $\cL_{2}$-valued and $\Phi\cap C^{\infty}_{0}(B_{n})$ is dense in $\cL_{2}(B_{n})$, we have that for $\omega\in\Omega''$, $t\in\bR_{+}$, and any $\phi\in C^{\infty}_{0}(B_{n})$ it holds that $$ (m_{t}^{\psi_{k}},\phi)= (m_{t}^{\psi_{j}},\phi) $$ as long as $i,j\geq n$. Then \eqref{9.3.7} implies that for any $\omega\in \Omega''$, $t\in\bR{_+}$, and $\phi\in C^{\infty}_{0}(B_{n})$ we have $(u_{t}^{\psi_{j}},\phi) =(u_{t}^{\psi_{k}},\phi)$ for all $j,k\geq n$. In particular, for any $\omega\in \Omega''$, $t \in\bR_{+}$, $n=1,2...$ it holds that $u_{t}^{\psi_{j}}=u_{t}^{\psi_{k}}$ as distributions on $B_{n}$ for $j,k\geq n$ and there exists a distribution $\bar{u}_{t}$ on $\bR^{d}$ such that $\bar{u}_{t}=u_{t}^{\psi_{k}}$ on $B_{n}$ for all $k\ge n$. Since $u_{t}^{\psi_{k}}=u_{t}\psi_{k}$ for almost all $(\omega,t)$, we have that $\bar{u}_{t}=u_{t}$ (as distributions on $\bR^{d}$) for almost all $(\omega,t)$. The inclusion $u\in\bW^{1}_{p}$ now yields $\bar{u}\in\bW^{1}_{p}$. It also follows from \eqref{9.3.7} that if $\omega\in\Omega''$, $t\in\bR_{+}$, and $\phi\in C^{\infty}_{0}$ is such that $\phi=0$ outside $B_{n}$, then for any $j\geq n$ $$ (\bar{u}_{t},\phi)=(u_{t}^{\psi_{j}},\phi) =\int_{0}^{t}(L_{s}u_{s}-\lambda u_{s}+D_{i}f^{i}_{s} +f^{0}_{s},\phi)\,ds+(m^{\psi_{j}}_{t},\phi). $$ By having in mind \eqref{1.26.1} we conclude that for any $\phi\in C^{\infty}_{0}$ with probability one for all $t\in\bR_{+}$ $$ (\bar{u}_{t},\phi) =(u_{0},\phi)+ \int_{0}^{t}(L_{s}u_{s}-\lambda u_{s}+D_{i}f^{i}_{s} +f^{0}_{s},\phi)\,ds $$ $$ +\sum_{k=1}^{\infty}\int_{0}^{t}(\Lambda^{k}_{s} u_{s}+g^{k}_{s},\phi)\,dw^{k}_{s}. $$ Now it only remains to observe that since $\bar{u}_{s}=u_{s}$ for almost all $(\omega,s)$, we can replace $u_{s}$ with $\bar{u}_{s}$ in the above equation. The theorem is proved. \mysection{It\^o's formula for the product of two processes of class $\cW^{1}_{2,loc}(\tau)$} \label{section 1.1.2} The results of this section will be used in a few places below, in particular, in the proof of Lemma \ref{lemma 10.8.2}. Recall that the spaces $\cW^{1}_{p}(\tau)$ are introduced in Definition \ref{definition 3.16.1}. \begin{theorem} \label{theorem 10.14.1} Let $\tau$ be a stopping time and let $u,\tilde{u}$, $f^{j}$, $\tilde{f}^{j}$, $g=(g^{1},g^{2},...)$, $\tilde{g}=(\tilde{g}^{1},\tilde{g}^{2},...)$ be some functions such that for any $\phi\in C^{\infty}_{0}$ we have $\phi u,\phi\tilde{u}\in\cW^{1}_{2}(\tau)$, $\phi f^{j},\phi\tilde{f}^{j}\in \bL_{2}(\tau)$, $j=0,...,d$, and $\phi g , \phi\tilde{g} \in \bL_{2}(\tau)$. Assume that in the sense of generalized functions $$ du_{t}=(D_{i}f^{i}_{t}+f^{0}_{t})\,dt+g^{k}_{t}\,dw^{k}_{t}, \quad d\tilde{u}_{t}=(D_{i}\tilde{f}^{i}_{t}+ \tilde{f}^{0}_{t})\,dt+\tilde{g}^{k}_{t}\,dw^{k}_{t}, \quad t\leq\tau. $$ Then $$ d(u_{t}\tilde{u}_{t})=\big[ \tilde{u}_{t}(D_{i}f^{i}_{t}+f^{0}_{t}) +u_{t}(D_{i}\tilde f^{i}_{t}+\tilde f^{0}_{t})+h_{t}\big]\,dt $$ $$ + (\tilde{u}_{t}g^{k}_{t}+u_{t}\tilde{g}^{k}_{t})\,dw^{k}_{t}, \quad t\leq\tau, $$ where $h_{t}:= (g _{t},\tilde{g} _{t} )_{\ell_{2}}$, in the sense of generalized functions, that is, for any $\phi\in C^{\infty}_{0}$, with probability one, $$ (u_{t\wedge\tau}\tilde{u}_{t\wedge\tau},\phi)= (u_{0}\tilde{u}_{0},\phi) +\int_{0}^{t}I_{s\leq\tau} (\tilde{u}_{s}g^{k}_{s}+u_{s}\tilde{g}^{k}_{s} ,\phi)\,dw^{k}_{s} $$ \begin{equation} \label{10.16.1} +\int_{0}^{t}I_{s\leq\tau} \big[(\tilde{u}_{s}f ^{0}_{s},\phi) -\big( f^{i}_{s},D_{i}(\tilde{u}_{s}\phi)\big)+ (u_{s}\tilde{f}^{0}_{s},\phi) -\big( \tilde{f}^{i}_{s},D_{i}(u _{s}\phi)\big) +(h_{s},\phi)\big]\,ds \end{equation} for all $t$. \end{theorem} Proof. To prove \eqref{10.16.1}, we only need to consider the case that $\tilde{u}=u$. Indeed, then by writing down the stochastic differential of $|u_{t}+\lambda\tilde{u}_{t}|^{2}$, where $\lambda$ is an arbitrary constant, and comparing the coefficients of $\lambda$, we would come to \eqref{10.16.1}. In other words, to prove \eqref{10.16.1}, we need only prove that for any $\phi \in C^{\infty}_{0}$ with probability one $$ (u^{2}_{t\wedge\tau} ,\phi)= (u^{2}_{0} ,\phi) +2\int_{0}^{t}I_{s\leq\tau} (u_{s}g^{k}_{s} ,\phi)\,dw^{k}_{s} $$ \begin{equation} \label{10.16.3} +\int_{0}^{t}I_{s\leq\tau} \big[2(u_{s}f ^{0}_{s},\phi) -2\big( f^{i}_{s},D_{i}(u_{s}\phi)\big) +(|g_{s}|^{2}_{\ell_{2}},\phi)\big]\,ds \end{equation} for all $t$. Next, observe that for any $\psi,\phi\in C^{\infty}_{0}$, with probability one $$ (\psi u_{t\wedge\tau},\phi)= (\psi u_{0},\phi)+ \int_{0}^{t}I_{t\leq\tau} (\psi g^{k}_{s},\phi)\,dw^{k}_{s} $$ $$ +\int_{0}^{t}I_{t\leq\tau}\big[ (\psi f^{0}_{s}-f^{i}_{s}D_{i}\psi,\phi) -(\psi f^{i}_{s},D_{i}\phi) \big]\,dt. $$ for all $t$. This means that $$ d(\psi u_{t})=(\psi f^{0}_{t}-f^{i}_{t}D_{i}\psi +D_{i}(\psi f^{i}_{t}))\,dt +\psi g^{k}_{t} \,dw^{k}_{t},\quad t\leq\tau. $$ By well-known results, in particular, by It\^o's formula (see, for instance \cite{Kr09_3}) there is a set $\Omega'\subset\Omega$ of full probability such that (i) $\psi u_{t\wedge\tau}I_{\Omega'}$ is a continuous $\cL_{2}$-valued $\cF_{t}$-adapted function on $[0,\infty)$; (ii) for all $t\in[0,\infty)$ and $\omega\in\Omega'$, It\^o's formula holds: $$ \int_{\bR^{d}}|\psi u_{t\wedge\tau}|^{2}\,dx =\int_{\bR^{d}}|\psi u_{0}|^{2}\,dx +2 \int_{0}^{t }I_{s\leq\tau} \int_{\bR^{d}} \psi^{2} u _{s} g^{k }_{s}\,dx\,dw^{k}_{s} $$ \begin{equation} \label{4.19.5} + \int_{0}^{t }I_{s\leq\tau} \big( \int_{\bR^{d}}\big[2 u_{ s}f^{0}_{ s}\psi^{2} -2 f^{i}_{ s} D_{i}(\psi^{2} u_{ s}) + \psi^{2}| g_{ s }|_{\ell_{2}}^{2} \big]\,dx\big)\,d s. \end{equation} This proves \eqref{10.16.3} if we replace there $\phi$ with $\psi^{2}$. However, for any $\phi\in C^{\infty}_{0}$ one can find $\psi_{1},\psi_{2}\in C^{\infty}_{0}$ such that $\phi=\psi_{1}^{2}-\psi_{2}^{2}$. Indeed, one can take sufficiently large $N,R>0$ and take $\psi_{1}(x)=\exp(-(R^{2}-|x|^{2})^{-1})$ for $|x|<R$ and $\psi_{1}(x)=0$ for $|x|\geq R$ and define $\psi_{2}=(\psi_{1}^{2}-\phi)^{1/2}$. This implies that \eqref{10.16.3} holds for any $\phi\in C^{\infty}_{0}$ with probability one for all $t$ and proves the theorem. \begin{corollary} \label{corollary 09.10.29.1} Let $u,f,g$ be as in Theorem \ref{theorem 10.14.1}, let a nonrandom $\psi\in W^{1}_{2}$, and let a random process $x_{t}$ be given as $$ x_{t}=\int_{0}^{t}\sigma^{k}_{s}\,dw^{k}_{s} +\int_{0}^{t}b_{s}\,ds $$ for some predictable $\bR^{d}$-valued functions $\sigma^{k}_{t}$ and $b_{t}$ such that $$ E\int_{0}^{\tau}\big(\sum_{k}|\sigma^{k}_{t}|^{2} +|b_{t}|\big)\,dt<\infty. $$ Then in the sense of generalized functions $$ d(u_{t}\psi_{t})=\big[D_{i}(u_{t}a^{ij}_{t}D_{j}\psi_{t}) -a^{ij}_{t}(D_{i}u_{t})D_{j}\psi_{t} +u_{t}b_{t}^{i}D_{i}\psi_{t}+D_{i}(\psi_{t}f^{i}_{t}) $$ $$ -f^{i}_{t}D_{i}\psi_{t}+\psi_{t}f^{0}_{t} +g^{k}_{t}\sigma^{ik}_{t}D_{i}\psi_{t}\big]\,dt +\big[\psi_{t}g^{k}_{t}+u_{t}\sigma^{ik}_{t}D_{i} \psi_{t}\big]\,dw^{k}_{t},\quad t\leq\tau, $$ where $\psi_{t}(x)=\psi(x+x_{t})$ and $2a^{ij}_{t} =\sigma^{ik}_{t}\sigma^{jk}_{t}$. \end{corollary} Indeed, observe that by It\^o's formula and the stochastic Fubini theorem, for any $\phi \in C^{\infty}_{0}$, $$ \int_{\bR^{d}}\psi_{t\wedge\tau}\phi\,dx= \int_{\bR^{d}}\psi(x) \phi(x-x_{t\wedge\tau})\,dx= \int_{\bR^{d}}\psi\phi\,dx $$ $$ +\int_{0}^{t}I_{s\leq\tau}\int_{\bR^{d}} \psi_{s}[a^{ij}_{s}D_{ij}\phi -b^{i}_{s}D_{i}\phi]\,dx\,ds+ \int_{0}^{t}I_{s\leq\tau}\int_{\bR^{d}}\sigma^{ik}_{s} \phi D_{i}\psi_{s}\,dx\,dw^{k}_{s}, $$ where the coefficient of $ds$ equals $$ \int_{\bR^{d}} \phi[a^{ij}_{s} D_{ij}\psi_{s}+ b^{i}_{s} D_{i}\psi_{s}]\,dx. $$ Furthermore, for instance, $$ E\int_{0}^{\tau}\int_{\bR^{d}}\sum_{k} |\sigma^{ik}_{s} D_{i}\psi_{s}|^{2}\,dx\,ds \leq E\int_{0}^{\tau}\int_{\bR^{d}}\sum_{k} |\sigma^{ k}_{s} |^{2}| D\psi_{s}|^{2}\,dx\,ds $$ $$ =\int_{\bR^{d}}| D\psi |^{2}\,dx\, E\int_{0}^{\tau}\sum_{k} |\sigma^{ k}_{s} |^{2}\,ds<\infty. $$ It follows that $\psi_{\cdot}\in\cW^{1}_{2}(\tau)$ and $$ d\psi_{t}=\big[D_{i}(a^{ij}_{t}D_{j}\psi_{t}) +b^{i}_{t}D_{i}\psi_{t}\big]\,dt +\sigma^{ik}_{t}D_{i}\psi_{t}\,dw^{k}_{t} $$ in the sense of generalized functions, so that the desired result follows from Theorem \ref{theorem 10.14.1}. \mysection{Kalman-Bucy filter} \label{section 11.8.1} We take a $T\in(0,\infty)$ and on $[0,T]$ consider a $d_{1}$-dimensional two component process $z_{t}=(x_{t},y_{t})$ with $x_{t}$ being $d$-dimensional and $y_{t}$ $(d_{1}-d)$-dimensional. We assume that $z_{t}$ is a diffusion process defined as a solution of the system \begin{equation}\begin{split} \label{eq3.2.14} & dx_{t}=b(t,z_{t}) dt+\theta (t,y_{t})dw_{t}, \\ & dy_{t}=B(t,z_{t}) dt+\Theta (t,y_{t})dw_{t} \end{split}\end{equation} with some initial data. \begin{assumption} \label{asm3.2.15} The functions $b$, $\theta $, $B$ and $\Theta $ are Borel measurable functions of $(t,z)$ and $(t,y)$ as appropriate and $\theta$ and $\Theta $ are bounded and satisfy the Lipschitz condition with respect to $y$ with a constant independent of $t$. We have $$ b(t,z)=x^{*}\dot{b}(t,y)+b(t,0,y),\quad B(t,z)=x^{*}\dot{B}(t,y)+B(t,0,y), $$ where $\dot{b}$ and $\dot{B}$ are bounded matrix-valued functions of appropriate dimensions, $b(t,0)$ and $B(t,0)$ are bounded, and $\dot{b}(t,y)$, $\dot{B}(t,y)$, $b(t,0,y)$, and $B(t,0,y)$ satisfy the Lipschitz condition with respect to $y$ with a constant independent of $t$. \end{assumption} In the rest of the article we use the notation $$ D_{i}=\frac{\partial}{\partial x^{i}},\quad D_{ij}=D_{i}D_{j} $$ only for $i,j=1,...,d$. \begin{remark} Note that \begin{equation} \label{9.19.1} \dot{b}^{ij}(t,y) =D_{i}b^{j} (t,z),\quad\dot{B}^{ij}(t,y) =D_{i}B^{j} (t,z). \end{equation} \end{remark} Set \begin{equation} \label{eq3.2.19.4} \check{\theta } (t,y) =\begin{pmatrix} \theta(t,y)\\ \Theta(t,y) \end{pmatrix},\quad \check{a} (t,y) =\frac{1}{2} \check{\theta }\check{\theta }^{*}(t,y),\quad \check{b} (t,z) =\begin{pmatrix} b(t,z) \\ B(t,z) \end{pmatrix}, \end{equation} \begin{equation} \label{eq3.2.19.1} \check{L} (t,z) = \check{a}^{ij} (t,y) \frac{\partial^{2} }{\partial z^{i} \partial z^{j}}+ \check{b}^{i} (t,z)\frac{\partial }{\partial z^{i}}, \end{equation} where $t\in[0,T]$, $z=(x,y)\in \bR^{d_{1}}$, and we use the summation convention over all ``reasonable'' values of repeated indices, so that the summation in (\ref{eq3.2.19.1}) is performed for $i,j=1,...,d_{1}$. Observe that \begin{equation} \label{9.29.4} dz_{t}=\check{\theta }(t,z_{t})\,dw_{t}+ \check{b}(t,z_{t})\,dt. \end{equation} \begin{assumption} \label{asm3.2.16} The symmetric matrix $\check{a} (t,y)$ is uniformly nondegenerate. In particular, the matrix $\Theta \Theta^{*}$ is invertible and $$ \Psi :=(\Theta \Theta^{*} )^{-\frac{1}{2}} $$ is a bounded function of $(t,y)$. \end{assumption} \begin{remark} \label{remark 9.19.1} It is well known (see, for instance, \cite{Kr_10}) that in light of Assumption \ref{asm3.2.16} the matrix $$ \hat{a}(t,y)=a(t,y)-\alpha(t,y) $$ is uniformly (with respect to $(t,y)$) nondegenerate, where $$ a =\frac{1}{2}\theta \theta^{*} , \quad \alpha=\tfrac{1}{2}\sigma\sigma^{*},\quad \sigma =\theta\Theta^{*}\Psi , $$ \end{remark} \begin{remark} \label{remark 1.16.1} Everywhere below we use the stipulation that if we are given a function $\xi(t,x,y)$, then we denote \begin{equation} \label{1.16.1} \xi_{t}=\xi_{t}(x)=\xi(t,x,y_{t}) \end{equation} unless it is explicitly specified otherwise. For instance, $ \Psi_{t}=\Psi(t,y_{t})$, $\Theta_{t}=\Theta(t,y_{t})$, $ \sigma_{t} =\theta_{t} \Theta^{*}_{t}\Psi_{t}$. \end{remark} Next we introduce a few more notation. Let (note the size and shape of $\sfB$) $$ \sfB=\Psi B,\quad \sfB_{t}(x) =\Psi_{t}B_{t}(x)=\Psi(t,y_{t}) B(t,x,y_{t}) $$ and set \begin{equation} \label{eq3.2.19.2} L_{t}( x) = a^{ij}_{t} D_{i}D_{j}+ b^{i}_{t}(x)D_{i}\,, \end{equation} $$ L^{*}_{t}( x)u_{t}(x) = D_{i}D_{j}( a^{ij}_{t} u_{t}(x) ) - D_{i}(b^{i}_{t}( x)u_{t}(x) ) $$ \begin{equation} \label{e3.2.19.2} =D_{j}\big( a^{ij}_{t} D_{i}u_{t}(x) -b^{j}_{t}(x)u_{t}(x) \big), \end{equation} \begin{equation} \label{1.27.8} \Lambda^{k }_{t}( x)u_{t}(x) = \sigma^{ik}_{t} D_{i}u_{t}(x)+\sfB^{k}_{t}( x)u_{t}(x), \end{equation} \begin{equation}\label{eq3.2.19.3} \Lambda^{k*}_{t}( x)u_{t}(x) =-\sigma^{ik}_{t} D_{i}u_{t}(x) + \sfB^{k}_{t}(x) u_{t}(x), \end{equation} where $t\in[0,T]$, $x\in \bR^{d}$, $k=1,...,d_{1}-d$, and as above we use the summation convention over all ``reasonable'' values of repeated indices, so that the summation in (\ref{eq3.2.19.2}), (\ref{e3.2.19.2}), \eqref{1.27.8}, and (\ref{eq3.2.19.3}) is performed for $i,j=1,...,d$ (whereas in (\ref{eq3.2.19.1}) for $i,j=1,...,d_{1}$). Finally, by $\cF_{t}^{y}$ we denote the completion of $\sigma \{ y_{s}:s\leq t\}$ with respect to $P,\cF$. \begin{assumption} \label{assumption 9.20.1} There exists an $\varepsilon>0$ and a function $Q(x)=Q(\omega,x)$ which is $\cF^{y}_{0}$-measurable in $\omega$, quadratic in $x$, and (i) For all $x\in\bR^{d}$ (and $\omega$) $$ \varepsilon^{-1}|x|^{2}\geq x^{i}x^{j}D_{ij}Q \geq\varepsilon |x|^{2}; $$ (ii) We have $\pi_{0}e^{Q}\in \text{\rm tr}\,\cW^{1}_{p}$, where $\pi_{0}$ is the conditional density of $x_{0}$ given~$y_{0}$. \end{assumption} Assumption \ref{assumption 9.20.1} is satisfied, for instance, in the classical setting of the Kalman-Bucy filter when $\pi_{0}$ is a Gaussian density. \begin{theorem} \label{theorem 11.8.1} There exists a process $\bar{\pi } $ on $[0,T]$ such that (i) $\bar{\pi }_{t}$ is $\cF^{y}_{t}$-adapted and, for any $r\in[1,p]$, with probability one $\bar{\pi}_{t}$ is a continuous $\cL_{r}$-valued process on $[0,T]$ and $\bar{\pi}_{0}=\pi_{0}$; (ii) There exists an increasing sequence of $\cF^{y}_{t}$-stopping times $\tau_{m}\leq T$ such that $P(\tau_{m}=T)\to1$ and $\bar{\pi} \in\bW^{1}_{p}(\tau_{m})$ for any $m$; (iii) In the sense of Definition \ref{definition 3.20.01} for any $m$ \begin{equation} \label{10.20.2} d\bar{\pi}_{t}=\Lambda^{k*}_{t}\bar{\pi}_{t} \,d\tilde{y}^{k}_{t}+L^{*}_{t}\bar{\pi}_{t}\,dt,\quad t\leq \tau_{m} , \end{equation} where $$ \tilde{y}^{k}_{t}=\int_{0}^{t} \Psi^{kr}_{s}\,dy^{r}_{s}. $$ Furthermore, for any $m$ and $\phi\in C^{\infty}_{0}$ we have $\bar{\pi}\phi\in\cW^{1}_{p}(\tau_{m})$; (iv) We have $\bar{\pi}_{t}\geq 0$ for all $t\in[0,T]$ (a.s.), \begin{equation} \label{1.27.07} 0<\int_{\bR^{d}} \bar{\pi }_{t}(x)\,dx=(\bar{\pi }_{t},1)<\infty \end{equation} for all $t\in[0,T]$ (a.s.), and for any $t\in[0,T]$ and real-valued, bounded or nonnegative, (Borel) measurable function $f$ given on $\bR^{d}$ \begin{equation} \label{10.13.3} E[f(x_{t})|\cF_{t}^{y}]= \frac{(\bar{\pi }_{t},f)} {(\bar{\pi }_{t},1)}\quad \text{(a.s.).} \end{equation} \end{theorem} \begin{remark} Equation \eqref{10.13.3} shows (by definition) that $$ \pi_{t}(x):=\frac{\bar{\pi }_{t}(x) } {(\bar{\pi }_{t},1)} $$ is a conditional density of distribution of $x_{t}$ given $y_{s},s\leq t$. Since, generally, $(\bar{\pi }_{t},1)\ne1$, one calls $\bar{\pi }_{t} $ an unnormalized conditional density of distribution of $x_{t}$ given $y_{s},s\leq t$. Thus, Theorem \ref{theorem 11.8.1} allows us to characterize the conditional density and being combined with Theorem \ref{theorem 1.12.1} allows us to obtain fine regularity properties of it. \end{remark} The following result is obtained by repeating what is said after Theorem~\ref{theorem 1.12.1} and taking into account that with probability one $\tau_{m}=T$ for all large $m$. \begin{theorem} \label{theorem 1.17.1} (ii) For any $\phi\in C^{\infty}_{0}$ the process $\bar{\pi }_{t }\phi$ is continuous on $[0,T]$ as an $\cL_{p}$-valued process (a.s.); (ii) If $p>2$ and we have two numbers $\alpha$ and $\beta$ such that $$ \frac{2}{p}<\alpha<\beta\leq1, $$ then for any $\phi\in C^{\infty}_{0}$ (a.s.) $$ \bar{\pi }\phi\in C^{\alpha/2-1/p}([0,T], H^{1-\beta}_{p}). $$ In particular, if $p>d+2$, then (a) for any $\varepsilon \in(0,\varepsilon_{0}]$, with $\varepsilon_{0}=1-(d+2)/p$, (a.s.) for any $t\in[0,T]$ we have $\bar{\pi }_{t}\phi\in C^{\varepsilon_{0}- \varepsilon}(\bR^{d})$ and the norm of $\bar{\pi }_{t}\phi$ in this space is bounded as a function of $t$; (b) for any $\varepsilon$ as in (a) (a.s.) for any $x\in\bR^{d}$ we have $\bar{\pi }_{\cdot}(x)\phi(x)\in C^{(\varepsilon_{0}- \varepsilon)/2}([0,T])$ and the norm of $\bar{\pi }_{\cdot}(x)\phi(x)$ in this space is bounded as a function of $x$. \end{theorem} In the general filtering theory equation \eqref{10.20.2} is known as Zakai's equation. From the point of view of the Sobolev space theory of SPDEs the most unpleasant feature of \eqref{10.20.2} in our particular case is the presence of $\sfB^{k}_{t}(x)\bar{\pi}_{t}$ in the stochastic term with $\sfB^{k}_{t}(x)$ which is unbounded in $x$. However, in the theory of linear PDEs it was observed that if an equation has a zeroth order term and we know a particular nonzero solution, then the ratio of the unknown function and this particular solution satisfies an equation without zeroth order term (cf. \eqref{9.29.01}). The way to find a particular solution of \eqref{10.20.2} is suggested by filtering theory. Imagine that $\check{b}$ is affine with respect to $z$ and $\check{\theta}$ is independent of $z$. Then as easy to see $z_{t}$ is a Gaussian process and hence the conditional density of $x_{t}$ given $y_{s}$, $s\leq t$, is Gaussian, that is, its logarithm is a quadratic function in $x$. Therefore, we were looking for a particular solution as $e^{-Q_{t}(x)}$, where $Q_{t}(x)$ is a quadratic function with respect to $x$, and finding the equation for $Q_{t}(x)$ (see \eqref{11.9.1}) was pretty straightforward. After we ``kill'' the zeroth-order term our equation falls into the scheme of Section \ref{section 2.15.1} even though it still has growing first order coefficients in the deterministic part of the equation. Finding $\bar{\pi}_{t}$ in the described way allows us to follow the scheme suggested in \cite{KZ} thus avoiding using filtering theory. However, we still encounter an additional difficulty that certain exponential martingales may not have moments of order $>1$, unlike the situation in \cite{KZ}, and, to prove that they are martingales indeed, we use the Liptser-Shiryaev theorem (see \cite{LS}). This way of proceeding was used by Liptser in \cite{Li} (see also \cite{LS}) while treating filtering problem for the so-called conditionally Gaussian processes. Finding a particular solution of \eqref{10.20.2} is based on the following lemma which is probably well known. We give its proof in the end of Section \ref{section 1.1.3} just for completeness. Set \begin{equation} \label{9.10.4} \dot{\sfB}_{t}=\dot{B}_{t}\Psi_{t}. \end{equation} \begin{lemma} \label{lemma 11.1.1} The following system of equations about $d\times d$-symmetric matrix-valued process $W_{t}$, $\bR^{d}$-valued process $V_{t}$, and real-valued process $U_{t}$ \begin{equation} \label{9.12.1} \frac{d}{dt}W_{t}= (\dot{\sfB}_{t}\sigma_{t}^{*} -\dot{b}_{t} )W_{t}+W^{*}_{t}(\sigma _{t}\dot{\sfB}_{t}^{*} -\dot{b}^{*}_{t}) -2W^{*}_{t}\hat{a}_{t}W_{t}+\dot{\sfB}_{t} \dot{\sfB}_{t} ^{*}, \end{equation} $$ dV_{t}=-(W_{t}\sigma_{t}+\dot{\sfB}_{t}) \,d\tilde{y}_{t} $$ \begin{equation} \label{9.12.2} +[(\dot\sfB_{t}\sigma^{*}_{t}-\dot{b}_{t} )V_{t} -2W_{t}\hat{a}_{t}V_{t} +W_{t}(\sigma_{t}\sfB_{t}(0)-b_{t}(0) ) +\dot{\sfB}_{t}\sfB_{t}(0) ]\,dt, \end{equation} $$ dU_{t}=-(V^{*}_{t}\sigma_{t}+\sfB^{*}_{t}(0)) \,d\tilde{y}_{t} $$ \begin{equation} \label{9.15.1} +[a^{ij}_{t}W^{ij}_{t}+ V^{*}_{t}(\sigma_{t}\sfB_{t}(0)-b_{t}(0) ) -V^{*}_{t}\hat{a}_{t}V _{t} +\tfrac{1}{2}|\sfB_{t}(0)|^{2}+\text{\rm tr}\,\dot{b}_{t}]\,dt, \end{equation} has a unique $\cF^{y}_{t}$-adapted solution with initial conditions $W_{0}^{ij}=D_{ij}Q$, $V_{0}^{i}=D_{i}Q(0)$, $U_{0}=Q(0) $. Furthermore, $\varepsilon_{1}^{-1}(\delta^{ij}) \geq W_{t}\geq\varepsilon_{1}(\delta^{ij})$ on $[0,T]$, where $\varepsilon_{1}>0$ is a constant independent of $\omega$ and $t$ (depending on $T$ among other things). \end{lemma} Observe that the coefficients in \eqref{9.15.1} are independent of $x$. \begin{remark} Set \begin{equation} \label{11.11.1} Q_{t}(x)=\tfrac{1}{2}W^{ij}_{t}x^{i}x^{j}+ V^{i}_{t}x^{i}+U_{t}. \end{equation} Then by using It\^o's formula one easily checks that for any $x\in\bR^{d}$ $$ dQ_{t}(x)=-(\sigma^{ik}_{t}D_{i}Q_{t}(x)+ \sfB^{k}_{t}(x) )\,d\tilde{y}^{k}_{t} +\big[ a^{ij}_{t}D_{ij}Q_{t}(x)+ D_{i}b^{i}_{t} $$ \begin{equation} \label{11.9.1} +(\sigma^{ik}_{t}\sfB^{k}_{t}(x)-b^{i}_{t}(x)) D_{i}Q_{t}(x) - \hat{a}^{ij}_{t} (D_{i}Q_{t}(x))D_{j}Q_{t}(x) +\tfrac{1}{2}|\sfB_{t}(x)|^{2}\big]\,dt \end{equation} and $\eta_{t}=e^{-Q_{t}}$ satisfies \begin{equation} \label{9.10.5} d\eta_{t}(x)= \Lambda^{r*}_{t}\eta_{t}(x)\,d\tilde{y}^{r}_{t} + L^{*}_{t} \eta_{t}(x) \,dt. \end{equation} By the way, $Q_{t}(x)$ is a unique $\cF^{y}_{t}$-adapted function depending quadratically on $x$, satisfying \eqref{11.9.1}, and such that $Q_{0}=Q$. Indeed, uniqueness follows from the fact that $D_{ij}Q_{t}$, $D_{i}Q_{t}(0)$, and $Q_{t}(0)$ are easily shown to satisfy \eqref{9.12.1}, \eqref{9.12.2}, and \eqref{9.15.1}, respectively. \end{remark} Our method also allows us to derive the classical equations for the Kalman-Bucy filter. \begin{theorem} \label{theorem 11.11.1} Replace requirement (ii) in Assumption \ref{assumption 9.20.1} with the assumption that $\pi_{0}=e^{-Q}$. Then for any $t$ (a.s.) we have $\pi_{t}(x)=C_{t}e^{-Q_{t}(x)}$, where $c_{t}$ is a normalizing process obtained from the condition that $$ C_{t}\int_{\bR^{d}}e^{-Q_{t}(x)}\,dx=1. $$ \end{theorem} This theorem is proved in Section \ref{section 1.1.4}. \begin{remark} \label{remark 1.17.1} After just completing the square and finding the stochastic differential of the remaining term we find that $$ Q_{t}(x)=\tfrac{1}{2}|W^{1/2}_{t}x+W^{-1/2}_{t}V_{t}|^{2} +\int_{0}^{t}(V^{*}_{s}W^{-1}_{s}\dot{\sfB}_{s} -\sfB^{*}_{s}(0))\,d\tilde{y}_{s} $$ \begin{equation} \label{11.1.3} +\tfrac{1}{2}\int_{0}^{t}|\dot{\sfB}^{*}_{s}W_{s}^{-1} V_{s}-\sfB_{s}(0)|^{2}\,ds+A_{t}, \end{equation} with a bounded on $\Omega\times [0,T]$ function $$ A_{t}:=\int_{0}^{t}[a^{ij}_{s}W^{ij}_{s}+\text{\rm tr}\, \dot{b}_{s} -\tfrac{1}{2}\|W^{1/2}_{s}\sigma_{s} +W^{-1/2}_{s}\dot{\sfB}_{s}\|^{2} ]\,ds, $$ where for a matrix $u$ we use the notation $\|u\|^{2} =\text{\rm tr}\, uu^{*}$. This shows that in the situation of Theorem \ref{theorem 11.11.1} $$ \bar{x}_{t}:=E(x_{t}\mid\cF^{y}_{t})= \int_{\bR^{d}}x\pi_{t}(x)\,dx =-W_{t}^{-1}V_{t}, $$ $$ \Sigma_{t}:=E\big((x_{t}-\bar{x}_{t})(x _{t}-\bar{x}_{t})^{*} \mid\cF^{y}_{t}\big)=W_{t}^{-1} $$ and allows one to derive the classical Kalman-Bucy equations for $\bar{x}_{t}$ and $\Sigma_{t}$ from \eqref{9.12.1} and \eqref{9.12.2}. \end{remark} \mysection{An auxiliary function} \label{section 1.1.3} The assumptions from Section \ref{section 11.8.1} are supposed to hold. Set $$ \hat{b}^{i}_{t}(x)= \sigma^{ik}_{t}\sfB^{k}_{t}(x)- 2\hat{a}^{ij}_{t} D_{j}Q_{t}(x). $$ \begin{theorem} \label{theorem 9.19.01} The equation \begin{equation} \label{9.29.01} d\hat{\pi}_{t}= -\sigma^{ik}_{t}D_{i}\hat{\pi}_{t}\, d\tilde{y}^{k}_{t} +\big[ a_{t}^{ij}D_{ij} \hat{\pi }_{t}-b^{i}_{t}D_{i}\hat{\pi}_{t} +\hat{b}^{i}_{t}D_{i}\hat{\pi}_{t} \big]\,dt ,\quad t\leq T , \end{equation} with initial data $\hat{\pi}_{0}=e^{Q}\pi_{0}$ has a unique solution in the sense of Definition~\ref{definition 3.20.01}. \end{theorem} This theorem is a direct consequence of Remark \ref{remark 9.29.2} and Theorem \ref{theorem 3.16.1} since the coefficients $b$ and $\hat{b}$ in \eqref{9.29.01} are affine functions of $x$ and have {\em bounded\/} derivatives in $x$. \begin{lemma} \label{lemma 10.30.1} Almost surely $\hat{\pi}_{t}$ is a continuous $\cL_{p}$-valued process on $[0,T]$. Furthermore, $G _{t}\|\hat{\pi}_{t}\|^{p}_{\cL_{p}}$ is a decreasing function of $t$ (a.s.), where $G_{t}$ is a bounded function on $\Omega\times[0,T]$ defined by $$ G_{t}:=\exp\int_{0}^{t}( D_{i}\hat{b}^{i}_{s}-D_{i}b^{i}_{s})\,ds =\exp\int_{0}^{t}\text{\rm tr}\,(\sigma_{s}\dot{\sfB}^{*}_{s} -\hat{a}_{s}W_{s}-\dot{b}_{s})\,ds. $$ In particular, on the set where $\tau:= T\wedge\inf\{t\geq0: \|\hat{\pi}_{t}\|_{\cL_{p}}=0\}<T$ we have $\|\hat{\pi}_{t}\|_{\cL_{p}}=0$ for $\tau\leq t\leq T$ (a.s.). \end{lemma} Proof. Set $$ \xi^{i}_{t}=\int_{0}^{t} \sigma^{ik}_{s}\,d\tilde{y}^{r}_{s},\quad\xi_{t}= (\xi^{i}_{t}),\quad \tau_{m}=T\wedge\inf\{t\geq0:|z_{t}|+|\xi_{t}|\geq m\}. $$ The purpose to stop $z_{t}$ is that on $\opar0,\tau_{m}\cbrk$ we have $$ |\sigma_{t}\sfB_{t}(x)|+|b_{t}(x)|+ |\hat{b}_{t}(x)|\leq N(1+|x|), $$ where the constant $N$ is independent of $\omega,t,x$. Why we also stop $\xi_{t}$ will become clear later. Observe that for any $\psi\in C^{\infty}_{0}$ the process $\psi\hat{\pi}_{t}$ satisfies an equation obtained by multiplying through \eqref{9.29.01} by $\psi$. Then after writing $\psi D_{i}(a^{ij}_{t} D_{j}\hat{\pi}_{t})$ as $ D_{i}(\psi a^{ij}_{t} D_{j}\hat{\pi}_{t})- a^{ij}_{t} (D_{j}\hat{\pi}_{t})D_{i}\psi$ and noting that the other coefficients multiplied by $\psi$ are bounded functions on $\opar0,\tau_{m}\cbrk\times\bR^{d}$ we see that \begin{equation} \label{11.1.1} \psi\hat{\pi}_{t}\in \cW^{1}_{p}(\tau_{m}) \end{equation} for any $m$. It follows from \cite{Kr09_3} that with probability one $\psi\hat{\pi}_{t\wedge\tau_{m}}$ is a continuous $\cL_{p}$-valued process and since, for each $\omega$, $\tau_{m}=T$ if $m$ is sufficiently large, with probability one $\psi\hat{\pi}_{t }$ is a continuous $\cL_{p}$-valued process on $[0,T]$ for any $\psi\in C^{\infty}_{0}$. Then take a nonnegative radially symmetric and radially decreasing function $\phi\in C^{\infty}_{0}$ such that $|D\phi|\leq1$, introduce $\phi^{n} (x) =\phi(x/n)$, $n=1,2,...$, $$ \phi^{n}_{t}(x)=\phi^{n} (x-\xi_{t}) $$ and use Corollary \ref{corollary 09.10.29.1} with $\tau_{m}$ in place of $\tau$ (recall \eqref{11.1.1}). Then we find $$ d(\hat{\pi}_{t}\phi^{n}_{t})=-\sigma^{ik}_{t}D_{i}( \hat{\pi}_{t}\phi^{n}_{t})\,d\tilde{y}^{k}_{t} +\phi^{n}_{t} ( \hat{b}^{i}_{t}-b^{i}_{t} ) D_{i}\hat{\pi}_{t} \,dt $$ \begin{equation} \label{10.29.5} +\big[ D_{i}(\phi^{n}_{t}a^{ij}_{t}D_{j}\hat{\pi}_{t})- (\hat{a}^{ij}_{t}+a^{ij}_{t})(D_{i}\phi^{n}_{t})D_{j}\hat{\pi}_{t} + D_{i}(\hat{\pi}_{t}\alpha^{ij}_{t}D_{j}\phi^{n}_{t}) \big]\,dt,\quad t\leq \tau_{m} . \end{equation} As above we conclude that $\phi^{n}_{t} \hat{\pi}\in\cW^{1}_{p}(\tau_{m})$ and that, owing to \cite{Kr09_3}, with probability one $\phi^{n}_{t}\hat{\pi}_{t}$ is a continuous $\cL_{p}$-valued process and (a.s.) $$ \|\phi^{n}_{t\wedge\tau_{m}} \hat{\pi}_{t\wedge\tau_{m}}\|_{\cL_{p}}^{p} =\|\phi^{n}\hat{\pi}_{0}\|_{\cL_{p}}^{p} + I^{1n}_{t}+I^{2n}_{t} +I^{3n}_{t} $$ for all $t$, where $$ I^{1n}_{t}=-p(p-1)\int_{0}^{t\wedge\tau_{m}} a^{ij}_{s}\int_{\bR^{d}} |\phi^{n}_{s}|^{p}|\hat{\pi}_{s}|^{p-2}(D_{i}\hat{\pi}_{s}) D_{j}\hat{\pi}_{s}\,dx\,ds\leq0, $$ $$ I^{2n}_{t}= -\int_{0}^{t\wedge\tau_{m}} \big[D_{i}\hat{b}^{i}_{s}- D_{i}b^{i}_{s}\big] \int_{\bR^{d}} |\phi^{n}_{s}|^{p}|\hat{\pi}_{s}|^{p} \,dx\,ds, $$ $$ I^{3n}_{t}=\int_{0}^{t\wedge\tau_{m}} \int_{\bR^{d}} |\hat{\pi}_{s}|^{p}\psi^{n}_{s}\,dx\,ds, $$ $$ \psi_{s}^{n}= pa^{ij}_{s}D_{ij}|\phi^{n}_{s} |^{p}+ (p-1)(p-2)|\phi^{n}_{s}|^{p-2}\alpha^{ij}_{s} (D_{i}\phi^{n}_{s})D_{j}\phi^{n}_{s} $$ $$ +(2-p)|\phi^{n}_{s}|^{p-1} \alpha^{ij}_{s}D_{ij}\phi^{n}_{s}\big] +(b^{i}_{s}-\hat{b}^{i}_{s})D_{i}|\phi^{n}_{s}|^{p} , $$ where for simplicity of notation the argument $x$ is dropped. Observe that $|D\phi^{n}_{s}| \leq 1/n$ and for $s\leq\tau_{m}$ we have $|b _{s}-\hat{b} _{s}|\leq N(1+|x|)$, where $N$ is independent of $s,x$, and $\omega$. Furthermore, $D\phi^{n}_{s}\to0$ as $n\to\infty$ and for $s<\tau_{m}$ $$ |x|\,|D \phi^{n}_{s}(x)|= \frac{|x|}{n}\big|(D \phi)\big(\frac{x-\xi_{s}}{n}\big)\big| \leq \frac{|\xi_{s}|}{n} +\frac{|x-\xi_{s}|}{n}\, \big|(D \phi)\big(\frac{x-\xi_{s}}{n}\big)\big| $$ $$ \leq m+\sup_{y}|y|\,|D\phi(y)|. $$ By adding that $\hat{\pi}\in\bW^{1}_{p}(T)$, we conclude that $I^{3n}_{t}\to 0$ uniformly in $t$ (a.s.). Analyzing $I^{1n}_{t}$ and $I^{2n}_{t}$ is almost trivial and $$\|\phi^{n}_{t\wedge\tau_{m}} \hat{\pi}_{t\wedge\tau_{m}}\|_{\cL_{p}}^{p} \to \| \hat{\pi}_{t\wedge\tau_{m}}\|_{\cL_{p}}^{p} $$ as $n\to\infty$ by the monotone convergence theorem. It follows that (a.s.) for all $t$ $$ \| \hat{\pi}_{t\wedge\tau_{m}}\|_{\cL_{p}}^{p} =\| \hat{\pi}_{0}\|_{\cL_{p}}^{p} -\int_{0}^{t\wedge\tau_{m}} (D_{i}\hat{b}^{i}_{s}- D_{i}b^{i}_{s}) \|\hat{\pi}_{s}\|_{\cL_{p}}^{p} \,ds $$ $$ -p(p-1)\int_{0}^{t\wedge\tau_{m}} a^{ij}_{s}\int_{\bR^{d}} |\hat{\pi}_{s}|^{p-2}(D_{i}\hat{\pi}_{s}) D_{j}\hat{\pi}_{s}\,dx\,ds . $$ Obviously on can drop $\tau_{m}$ in this formula and then obtain that (a.s.) for all $t\leq T$ $$ G_{t} \| \hat{\pi}_{t }\|_{\cL_{p}}^{p} =\| \hat{\pi}_{0}\|_{\cL_{p}}^{p} -p(p-1)\int_{0}^{t }G _{s} a^{ij}_{s}\int_{\bR^{d}} |\hat{\pi}_{s}|^{p-2}(D_{i}\hat{\pi}_{s}) D_{j}\hat{\pi}_{s}\,dx\,ds, $$ which implies that $G_{t} \| \hat{\pi}_{t }\|_{\cL_{p}}^{p}$ is decreasing and continuous (a.s.). Furthermore, since $\phi^{n}_{t}\hat{\pi}_{t}$ are continuous $\cL_{p}$-valued processes, $\hat{\pi}_{t}$ is at least a weakly continuous $\cL_{p}$-valued function, but since $\|\hat{\pi}_{t}\|^{p}_{\cL_{p}}$ is (absolutely) continuous, $\hat{\pi}_{t}$ is strongly continuous. This proves the lemma. \begin{remark} After we know that $\hat{\pi}_{t}$ is a continuous $\cL_{p}$-valued process on $[0,T]$ the last assertion of Lemma \ref{lemma 10.30.1} can be also obtained from uniqueness of solutions of \eqref{9.29.01} because the $\tau $ in Lemma \ref{lemma 10.30.1} is a stopping time and $\hat{\pi}_{t\wedge\tau}$ is obviously a solution of \eqref{9.29.01} implying that on the set where $\tau<T$ we have $\hat{\pi}_{t}=0$ for $\tau\leq t\leq T$. \end{remark} Before stating the following lemma we introduce a stipulation accepted throughout the rest of the paper that if we are given a function $\xi(t,x,y)$, then we denote \begin{equation} \label{10.10.1} \tilde{\xi}_{t}=\xi_{t}(x_{t}) =\xi(t,x_{t},y_{t}). \end{equation} The reader encountered above already one of these abbreviated notation (see \eqref{1.16.1}). \begin{lemma} \label{lemma 10.29.1} Introduce $$ \tilde{w} _{t}=\int_{0}^{t}\Psi _{s}\Theta_{s}\,dw _{s}, \quad \tilde{\sfB}_{t}=\sfB_{t}(x_{t})=\Psi(t,y_{t})B(t,x_{t},y_{t}). $$ Then $\tilde{w} _{t}$ is a Wiener process and the process $$ \rho_{t}=\rho_{t}(\tilde{\sfB},d\tilde{w}) =\exp(-\int_{0}^{t} \tilde{\sfB}^{k}_{s}\,d\tilde{w}^{k}_{s} -\tfrac{1}{2}\int_{0}^{t}|\tilde{\sfB}_{s}|_{\ell_{2}}^{2}\,ds) $$ is a martingale on $[0,T]$. \end{lemma} Proof. The first assertion follows from L\'evy's theorem. To prove the second one observe that $$ \int_{0}^{t} \tilde{\sfB}^{k}_{s}\,d\tilde{w}^{k}_{s}= \int_{0}^{t} \tilde{\sfB}^{*}_{s}\Psi _{s}\Theta_{s}\,dw _{s}. $$ Furthermore, the system $$ dx_{t}=\big(b(t,z_{t})-\theta(t,y_{t}) \Theta^{*}(t,y_{t})\Psi^{2}(t,y_{t}) B(t,z_{t})\big)\,dt+\theta(t,y_{t})\,dw_{t}, $$ $$ dy_{t}=\Theta(t,y_{t})\,dw_{t}, $$ which is obtained from \eqref{eq3.2.14} by formal application of the measure change, has a unique solution with initial data $z_{0}$ since its coefficients are locally Lipschitz in $z$ and grow as $|z|\to\infty$ not faster than linearly. In this situation by the Liptser-Shiryaev theorem $\rho$ is a martingale since $$ \int_{0}^{T}| \Psi (t,y(t))B(t,x(t),y(t))|^{2}\,dt<\infty $$ for any deterministic functions $x(t)$ and $y(t)$ which are continuous on $[0,T]$. The lemma is proved. \begin{lemma} \label{lemma 10.8.1} The process $\hat{\pi}_{t}$ is $\cF^{y}_{t}$-adapted. \end{lemma} Proof. Observe that in equation \eqref{9.29.01} we have $$ d\tilde{y}^{k}_{t}=\Psi^{kr}_{t}\,dy^{r}_{t}=d\tilde{w}^{k} _{t} +\tilde{\sfB}^{k}_{t}\,dt, $$ where, as it is pointed out above, $\tilde{w} _{t}$ is a Wiener process. Furthermore, the processes $\tilde{y}^{k}_{t}$ is $\cF^{y}_{t}$-adapted since such are $\Psi^{kr}_{t}$ and equation \eqref{9.29.01} is rewritten as $$ d\hat{\pi}_{t}= -\sigma^{ik}_{t}D_{i}\hat{\pi}_{t}\,d\tilde{w}^{k}_{t} +\big[ D_{i}( a^{ij}_{t} D_{ j} \hat{\pi }_{t})-b^{i}_{t}D_{i} \hat{\pi}_{t} $$ \begin{equation} \label{10.7.1} +( \hat{b}^{i}_{t}-\sigma^{ik}_{t}\tilde{\sfB}^{k}_{t} )D_{i}\hat{\pi}_{t} \big]\,dt ,\quad t\leq T . \end{equation} Here $\sigma^{ik}_{t}\tilde{\sfB}^{k}_{t}$ is independent of $x$ and for each $\omega$ the trajectories of $\sigma^{ik}_{t}\tilde{\sfB}^{k}_{t}$ are locally bounded on $\bR_{+}$, which shows that in order to be able to apply Theorem \ref{theorem 10.6.1} it only remains to refer to Lemma \ref{lemma 10.29.1}. The lemma is proved. \begin{lemma} \label{lemma 10.8.2} The assertions (i)-(iii) of Theorem \ref{theorem 11.8.1} hold for $\bar{\pi}_{t}:=e^{-Q_{t}}\hat{\pi}_{t}$. \end{lemma} Proof. Assertion (i) of Theorem \ref{theorem 11.8.1} follows immediately from Lemma \ref{lemma 10.30.1}, the continuity of $Q_{t}$, and the boundedness of $W_{t}=(D_{ij}Q_{t})$ away from zero. To prove assertion (ii) notice that $\hat{\pi}\in\bW^{1}_{p}(T)$ and $$ \int_{0}^{t}\|\bar{\pi}_{s}\|_{W^{1}_{p}}^{p}\,ds $$ is an $\cF^{y}_{t}$-adapted continuous process on $[0,T]$. Then after introducing $$ \tau'_{m}=T\wedge\inf\{t\geq0: \int_{0}^{t}\|\bar{\pi}_{s}\|_{W^{1}_{p}}^{p}\,ds\geq m\} $$ we get that $\bar{\pi}\in\bW^{1}_{p}(\tau'_{m})$ and $\tau'_{m}=T$ for all large $m$ (a.s.). We now prove that $\bar{\pi}$ satisfied \eqref{10.20.2} define $\Phi_{t}=\Psi_{t}^{-1}$ and observe that $$ (d\tilde{y}^{k}_{t})d\tilde{y}^{r}_{t}=\delta^{kr}\,dt,\quad dy^{k}_{t}=\Phi^{kr}_{t}\,d\tilde{w}^{r}_{t}+\tilde{B}^{k}_{t}\,dt, $$ ( $\tilde{B}_{t}=B(t,z_{t})$). Recall that $\eta_{t}(x)=\exp(-Q_{t}(x))$ satisfies equation \eqref{9.10.5} for each $x$ with probability one for all $t\in[0,T]$. It turns out that this equation also holds in the sense of generalized functions. Owing to the special structure of $Q_{t}$, this follows from the stochastic version of Fubini's theorem (see, for instance, Lemma 2.7 of \cite{Kr09_4}). Next, for $m=1,2,...$ set \begin{equation} \label{11.2.1} \tau''_{m}=T\wedge\inf\{t\geq0:|z_{t}|+|DQ_{t}(0)|\geq m\}. \end{equation} Note that for a constant $N_{0}$ independent of $m$ for $t<\tau''_{m}$ we have $$ |\sfB_{t}(x)| +|b_{t}(x)|\leq N_{0}(1+|x|+m), \quad|\tilde{\sfB}_{t}|+|\tilde{B}_{t}| \leq N_{0}(1+2m). $$ Furthermore, $D_{i}Q_{t}(x)=x^{j}D_{ij}Q_{t}+D_{i}Q_{t}(0)$, so that increasing $N_{0}$ if needed we may assume that for $t<\tau''_{m}$ $$ |DQ_{t}(x)|\leq N_{0}(1+ |x|+m). $$ Then as is easy to see (cf. \eqref{11.1.1}) $u_{t}:=\hat{\pi}_{t}$ and $\tilde{u}_{t}:=\eta_{t}$ satisfy the condition of Theorem~\ref{theorem 10.14.1} with appropriate $f,\tilde{f},g,\tilde{g}$ and $\tau''_{m}$ in place of $\tau$. By Theorem \ref{theorem 10.14.1} in the sense of generalized functions $$ d(\eta_{t}\hat{\pi}_{t}) =I^{r}_{t}\, d\tilde{y}^{r}_{t}+J_{t}\,dt,\quad t\leq\tau''_{m}, $$ where $$ I^{r}_{t}=\hat{\pi}_{t}\Lambda^{r*}_{t}\eta_{t} -\eta_{t} \sigma^{ir}_{t}D_{i}\hat{\pi}_{t} =\Lambda^{r*}_{t}(\eta_{t} \hat{\pi}_{t}), $$ $$ J_{t}=-(\eta_{t} \sfB^{k}_{t}- \sigma^{ik}_{t}D_{i}\eta_{t} )\sigma^{jk}_{t}D_{j}\hat{\pi}_{t} +\hat{\pi}_{t} L^{*}_{t} \eta_{t} $$ $$ +\eta_{t}\big[ a^{ij} _{t}D_{ij} \hat{\pi }_{t}-b^{i}_{t}D_{i}\hat{\pi}_{t} +( \sigma^{ik}_{t}\sfB^{k}_{t}+ 2\eta^{-1}\hat{a}^{ij}_{t} D_{j}\eta_{t})D_{i}\hat{\pi}_{t} \big] $$ $$ =\hat{\pi}_{t} L^{*}_{t} \eta_{t} +\eta_{t}(a^{ij} _{t}D_{ij} \hat{\pi }_{t}-b^{i}_{t}D_{i}\hat{\pi}_{t}) +2a^{ij}_{t}(D_{i}\hat{\pi}_{t}) D_{j}\eta_{t} =L^{*}_{t}(\eta_{t}\hat{\pi}_{t}). $$ In other words (see Theorem \ref{theorem 10.14.1}) for any $\phi\in C^{\infty}_{0}$ with probability one $$ (\bar{\pi}_{t\wedge\tau''_{m}},\phi) =(\bar{\pi}_{ 0},\phi) +\int_{0}^{t}I_{s\leq\tau''_{m}}(\bar{\pi}_{s},\Lambda^{k} _{s}\phi)\,d\tilde{y}^{k}_{s} +\int_{0}^{t}I_{s\leq\tau''_{m}}(\bar{\pi}_{s}, L_{s} \phi)\,ds $$ for all $t\geq0$. Obviously, one can take here $\tau_{m}: =\tau'_{m}\wedge\tau''_{m}$ in place of $\tau''_{m}$ and then after recalling that $\bar{\pi}\in\bW^{1}_{p} (\tau'_{m})$ one concludes that $\bar{\pi}$ is a solution of \eqref{10.20.2} in the sense of Definition \ref{definition 3.20.01}. The final assertion in (iii) is obtained in the same way as \eqref{11.1.1}. The lemma is proved. To better orient the reader it is worth noting that in the next lemma the second factor on the left in \eqref{11.5.1} contains the negative of two terms in \eqref{11.1.3}. \begin{lemma} \label{lemma 11.5.1} We have $$ \rho_{t}(\tilde{\sfB},d\tilde{w}) \exp\big(-\int_{0}^{t}(V^{*}_{s}W^{-1}_{s}\dot{\sfB}_{s} -\sfB^{*}_{s}(0))\,d\tilde{y}_{s} -\tfrac{1}{2}\int_{0}^{t}|\dot{\sfB}^{*}_{s}W_{s}^{-1} V_{s}-\sfB_{s}(0)|^{2}\,ds\big) $$ \begin{equation} \label{11.5.1} =\rho_{t}(\tilde{\sfB}- \sfB_{\cdot}(0)+ \dot{\sfB}^{*}W^{-1}V,d\tilde{w}). \end{equation} Furthermore, the right-hand side is a martingale on $[0,T]$. \end{lemma} Proof. The equality is obtained by simple manipulations. As in the proof of Lemma \ref{lemma 10.29.1}, to prove that \eqref{11.5.1} is a martingale we are going to use the Liptser-Shiryaev theorem by considering the system consisting of \eqref{9.29.4}, \eqref{9.12.1}, and \eqref{9.12.2}. We do not include \eqref{9.15.1} because $U_{t}$ does not enter \eqref{11.5.1}. First of all we find a smooth bounded, uniformly nondegenerate $d \times d$-matrix-valued function $F(W )$ such that $F(W_{t})=W_{t}$. The fact that this is possible follows from Lemma \ref{lemma 11.1.1}. Then set $$ A(t,z,W,V)=\Theta^{*}(t,y) \Psi^{2}(t,y ) \big(B(t,z )-B(t,0,y )+\dot{B}^{*}(t,y)F^{-1}(W ) V \big). $$ After changing the probability measure formally we arrive at the system consisting of \eqref{9.12.1} with $\sigma_{t}=\sigma(t,y_{t})$, $\hat{a}_{t} =\hat{a}(t,y_{t})$, and with $F(W_{t})$ in place of $W_{t}$ on the right and the following two equations $$ dz_{t}=\check{\theta}(t,y_{t})\,dw_{t} +\big[\check{b}(t,z_{t})-\check{\theta}(t,y_{t}) A(t,z_{t},W_{t},V_{t})\big]\,dt , $$ $$ dV_{t}=-\big(F(W_{t})\sigma(t,y_{t})+\dot{B}(t,y_{t}) \Psi(t,y_{t})\big)\Psi(t,y_{t}) \Theta(t,y_{t})\,dw_{t} $$ $$ +\big(F(W_{t})\sigma(t,y_{t})+\dot{B}(t,y_{t}) \Psi(t,y_{t})\big)\Psi(t,y_{t}) \Theta(t,y_{t})A(t,z_{t},W_{t},V_{t})\,dt $$ $$ -\big(F(W_{t})\sigma(t,y_{t})+\dot{B}(t,y_{t}) \Psi(t,y_{t})\big)\Psi(t,y_{t})B(t,z_{t})\,dt $$ $$ +\big[\big(\dot{B}(t,y_{t}) \Psi(t,y_{t})\sigma^{*}(t,y_{t})-\dot{b}(t,y_{t})\big)V_{t} -2F(W_{t})\hat{a}(t,y_{t})V_{t}\big]\,dt $$ $$ +\big[F(W_{t})\big(\sigma(t,y_{t})\Psi(t,y_{t})B(t,0,y_{t})- b (t,0,y_{t}) \big) +\dot{B}(t,y_{t}) \Psi^{2}(t,y_{t})B (t,0,y_{t}) \big]\,dt. $$ This system has a unique solution with prescribed initial data since its coefficients are locally Lipschitz continuous and may grow to infinity as $|z|+|W|+|V|\to\infty$ not faster than linearly. Moreover, $$ \int_{0}^{T}|A(t,z(t),W(t),V(t))|^{2}\,dt<\infty $$ for any functions $z(t),W(t),V(t)$ which are continuous on $[0,T]$. This implies that process \eqref{11.5.1} is a martingale on $[0,T]$ and the lemma is proved. {\bf Proof of Lemma \ref{lemma 11.1.1}}. Notice that \eqref{9.15.1} yields $U_{t}$ once $W_{t}$ and $V_{t}$ are found. Equation \eqref{9.12.2} is linear with respect to $V_{t}$ and proving the existence and uniqueness of its solution presents no difficulty if $W_{t}$ is known. Equation \eqref{9.12.1} can be considered for each $\omega$ separately. Then the theory of ODEs allows us to conclude that a unique solution exists until it blows up and it is $\cF^{y}_{t}$-adapted. Uniqueness implies that $W_{t}=W^{*}_{t}$. Furthermore, at least on a small time interval $W_{t}>0$. It turns out that $W_{t}>0$ on any interval of time where $W_{t}$ is bounded. Indeed, if not, then for some $t_{0}>0 $ we would have that $\det W_{t_{0}}=0$, $W_{t}$ is bounded on $[0,t_{0}]$ and $\det W_{t}>0$ for $t<t_{0}$. However, for $t<t_{0}$ \begin{equation} \label{11.8.1} \frac{d}{dt}\,\det W_{t }=\text{\rm tr}\, \dot{W}_{t}W^{-1}_{t}\det W_{t }, \end{equation} and $$ \text{\rm tr}\, \dot{W}_{t}W^{-1}_{t}=2\text{\rm tr}\,(\dot{\sfB}_{t}\sigma_{t}^{*} -\dot{b}_{t} )-2\text{\rm tr}\, \hat{a}_{t}W_{t}+\text{\rm tr}\,\dot{\sfB}_{t}\dot{\sfB} ^{*}_{t}W^{-1}_{t}, $$ where the last term is nonnegative as the trace of the product of two symmetric nonnegative matrices. It follows, that $\text{\rm tr}\, \dot{W}_{t}W^{-1}_{t}$ is bounded from below on $[0,t_{0})$ and hence equation \eqref{11.8.1} implies that $\det W_{t_{0}} >0$. Next, it turns out that the solution does not blow up on $[0,T]$. Indeed $$ \frac{d}{dt}\,\text{\rm tr}\, W_{t}W _{t}=4\text{\rm tr}\,(\dot{\sfB}_{t}\sigma_{t}^{*} -\dot{b}_{t} )W_{t}W _{t}+ 2\text{\rm tr}\,\dot{\sfB}_{t} \dot{\sfB}_{t}^{*}W_{t}-4\text{\rm tr}\, \hat{a}_{t} W^{3}_{t} , $$ where the last trace is nonnegative again on the interval of existence of $W_{t}$. Here $$ \text{\rm tr}\,\dot{\sfB}_{t} \dot{\sfB}_{t}^{*}W_{t}\leq N(\text{\rm tr}\, W^{2}_{t})^{1/2} \leq N+\text{\rm tr}\, W^{2}_{t}, $$ where $N$ is a constant. Also for two matrices $A$ and $W$ such that $W$ is symmetric and nonnegative it holds that $$ (\text{\rm tr}\, AW^{2})^{2}\leq\|A\|\,\|W^{2}\|\leq \|A\|(\text{\rm tr}\, W^{2})^{2}. $$ This and Gronwall's inequality imply that $W_{t}$ is bounded on $[0,T]$. Obviously the bound of $W_{t}$ is uniform with respect to $\omega$. The lower bound is also uniform since by the above $\det W_{t}$ is bounded away from zero on $[0,T]$ uniformly with respect to $\omega$. The lemma is proved. \mysection{Proof of Theorems \protect\ref{theorem 11.8.1} and \protect\ref{theorem 11.11.1}} \label{section 1.1.4} Take a function $\varphi\in C^{\infty}_{0}(\bR^{d_{1}})$ and let $c (t,y)$ be a smooth, bounded, and nonnegative function on $[0,T]\times\bR^{d_{1}-d}$. Recall that the operator $\check{L}$ is introduced in \eqref{eq3.2.19.1} and consider the following deterministic problem $$ \partial_{t}v (t,z)+ \check{L} v (t,z)-c (t,y)v (t,z) =0,\quad t\in[0,T],z\in\bR^{d_{1}}, $$ \begin{equation} \label{10.13.1} v (T,z) =\varphi(z),\quad z\in\bR^{d_{1}}. \end{equation} \begin{remark} \label{remark 9.20.1} By Theorem 2.5 of \cite{KP}, for any $\alpha\in(0,1)$ there exists a unique classical solution $v $ of \eqref{10.13.1} such that, for any $t\in[0,T]$, $v(t,\cdot)\in C^{2+\alpha}(\bR^{d_{1}})$ and the standard $C^{2+\alpha}(\bR^{d_{1}})$-norms of $v(t,\cdot)$ are bounded on $[0,T]$. If we denote by $z_{t}(s,z)$, $t\geq s$, the solution of system \eqref{eq3.2.14} which starts at $z$ at moment $s\leq T$, then by It\^o's formula we have $$ v (s, z)=E\varphi(z_{T }(s,z))\exp(-\int_{s}^{T } c_{r}( y_{r}(s,z))\,dr), $$ $$ |v(s,z)|\leq \sup|\varphi|P\{\tau(s,z)\leq T\} \leq \sup|\varphi|e^{N_{0}T}E e^{-N_{0}\tau(s,z)}, $$ where $N_{0}>0$ is an arbitrary constant, $\tau(s,z)$ is the first time $z_{t}(s,z)$ hits $\{z:|z|\leq R\}$, and $R$ is such that $\varphi(z)=0$ for $|z|\geq R$. Take an $m\geq0$ and introduce $\psi(z)=(1+|z|^{2})^{-m}$. It is not hard to see that, if $N_{0}$ is sufficiently large, then $$ \check{L}_{t}\psi(z)-N_{0}\psi(z) \leq0. $$ By It\^o's formula, for $|z|\geq R$, $$ \psi(R)E e^{-N_{0}\tau(s,z)}\leq \psi(z), $$ implying that for any $m\geq0$ there is a constant $N$ such that for all $(s,z)$ $$ |v(s, z)|\leq\frac{N}{(1+|z|^{2})^{m}}. $$ The argument in the proof of Lemma 4.11 of \cite{KZ} proves that the same estimate holds for $\partial v(s,z)/ \partial z^{i}$ and $\partial^{2} v(s,z)/ \partial z^{i}\partial z^{j}$, $i,j=1,...,d_{1}$. \end{remark} Before we come to a crucial point we state the following. \begin{lemma} \label{lemma 10.29.2} Let $\xi_{t}$ be a nonnegative continuous martingale on $[0,T]$ and let $\zeta_{t}$ be a continuous $\cF_{t}$-adapted process given on $[0,T]$ such that $\xi_{t}\zeta_{t}$ is a local martingale on $[0,T)$. Assume that $$ E\xi_{T}\sup_{[0,T]}|\zeta_{t}|<\infty. $$ Then $\xi_{t}\zeta_{t}$ is a martingale on $[0,T]$. \end{lemma} Proof. We need to prove that for any stopping time $\tau\leq T$ we have $E\xi_{\tau}\zeta_{\tau}=E\xi_{0}\zeta_{0}$. Here the left hand side equals $E\xi_{T}\zeta_{\tau}$ and we are given that there exists a sequence of stopping times $\tau_{n}\uparrow T$ such that $E\xi_{T}\zeta_{\tau \wedge\tau_{n}}=E\xi_{0}\zeta_{0}$. Using the dominated convergence theorem yields the desired result and proves the lemma. \begin{lemma} \label{lemma 10.8.02} The process $$ \rho_{t}e^{-\int_{0}^{t}c_{s}(y_{s})\,ds}\int_{\bR^{d}} v (t,x,y_{t})\bar{\pi}_{t}(x)\,dx $$ is a martingale on $[0,T]$. \end {lemma} Proof. Define ($c_{t}=c (t,y_{t})$, $v_{t}(x)=v(t,x,y_{t})$) $$ D^{y}_{k}=\frac{\partial}{\partial y^{k}},\quad D^{y}_{kr}=D^{y}_{k}D^{y}_{r},\quad C_{t}=\exp( -\int_{0}^{t}c_{s}\,ds),\quad \chi_{t}=C_{t}v_{t}\bar{\pi}_{t}. $$ We need to show that \begin{equation} \label{11.11.6} \rho_{t} \int_{\bR^{d}} \chi_{t}(x)\,dx \end{equation} is a martingale. Observe that by It\^o's formula and \eqref{10.13.1} we have $$ d [v_{t}(x )C_{t}]=d [v_{t} C_{t}]=C_{t}\big[ D^{y}_{k}v_{t} \,dy^{k}_{t} + \big(\partial_{t}v_{t} -c_{t}v_{t} +\check{a}^{kr}_{t} D^{y}_{kr} v_{t} \big)\,dt\big] $$ \begin{equation} \label{10.24.2} =C_{t}\big[D^{y}_{k}v_{t} \Phi^{kr}_{t}\,d\tilde{w}^{r}_{t}-\big(L_{t} v_{t} +2 \check{a}^{ik}_{t} D _{i}D^{y}_{k} v_{t}+ (B^{k}_{t}-\tilde{B}^{k}_{t}) D^{y}_{k}v_{t} \big)\,dt \big], \end{equation} where we dropped the arguments $x$ for shortness and, of course, $D^{y}_{k}v_{t}= (D^{y}_{k}v) (t,x,y_{t})$, $D^{y}_{kr}v_{t}= (D^{y}_{kr}v) (t,x,y_{t})$, and $D_{i}D^{y}_{k}v_{t}= (D_{i}D^{y}_{k}v) (t,x,y_{t})$. By the way, observe that $$ \sigma^{ir}_{t}\Phi^{kr}_{t} =2\check{a}^{ik}_{t},\quad B^{k}_{t}=\Phi^{kr}_{t} \sfB^{r}_{t}. $$ Similarly to the proof of Lemma \ref{lemma 10.8.2} we conclude that \eqref{10.24.2} holds in the sense of distributions and that Theorem \ref{theorem 10.14.1} is applicable to $v_{t}\bar{\pi}_{t}$ on the time interval $t\leq\tau_{m}$ for any $n$, where $\tau_{m}$ are taken from Lemma \ref{lemma 10.8.2}. It follows that for any $m$ for $t\leq\tau_{m}$ $$ d\chi_{t}= C_{t}(\bar{\pi}_{t}\Phi^{kr}_{t}D^{y}_{k}v_{t} +v_{t}\Lambda^{r*}_{t}\bar{\pi}_{t}) \,d\tilde{w}^{r}_{t} $$ $$ -C_{t}\bar{\pi}_{t}\big(L_{t} v_{t} + \sigma^{ir}_{t}\Phi^{kr}_{t} D _{i}D^{y}_{k} v_{t}+\Phi^{kr}_{t} (\sfB^{r}_{t}-\tilde{\sfB}^{r}_{t}) D^{y}_{k}v_{t} \big)\,dt $$ $$ +C_{t}v_{t}(L^{*}_{t}\bar{\pi}_{t}+ \tilde{\sfB}_{t}^{k} \Lambda^{k*}_{t}\bar{\pi}_{t})\,dt +C_{t}(D^{y}_{k}v_{t}) \Phi^{kr}_{t}\Lambda^{r*}_{t}\bar{\pi}_{t}\,dt. $$ It is convenient to rearrange the above terms by using the notation $$ \zeta^{r}_{t}=C_{t}(\bar{\pi}_{t}\Phi^{kr}_{t}D^{y}_{k}v_{t} +v_{t}\Lambda^{r*}_{t}\bar{\pi}_{t}). $$ We have $$ d\chi_{t}=\zeta_{t}^{r}\,d\tilde{w}^{r}_{t} +(\tilde{\sfB}^{r}_{t}\zeta^{r}_{t}+I^{1}_{t}+I^{2}_{t})\,dt, \quad t\leq\tau_{m}, $$ where $$ I^{1}_{t}= C_{t} (v_{t}L^{*}_{t}\bar{\pi}_{t} -\bar{\pi}_{t}L_{t} v_{t}),\quad I^{2}_{t}=-C_{t}\Phi_{t}^{kr}\sigma^{ir}_{t}( \bar{\pi}_{t} \sigma^{ir}_{t}D_{i}D^{y}_{k}v_{t}+ (D^{y}_{k}v_{t})D_{i}\bar{\pi}_{t} ) $$ $$ =-C_{t}\Phi_{t}^{kr}\sigma^{ir}_{t}D_{i}(\bar{\pi}_{t} D^{y}_{k}v_{t}). $$ In the integral form this means that for any $\phi \in C^{\infty}_{0}$ with probability one $$ (\chi_{t\wedge\tau_{m}},\phi)=(\chi_{0},\phi) +\int_{0}^{t}I_{s\leq\tau_{m}} (\zeta^{r}_{s},\phi)\, d\tilde{w}_{s}^{r}+ \int_{0}^{t}I_{s\leq\tau_{m}} \tilde{\sfB}^{r}_{s}(\zeta^{r}_{s},\phi)\,ds $$ $$ +\int_{0}^{t}I_{s\leq\tau_{m}}C_{s}a^{ij}_{s} (\bar{\pi}_{s}D_{j}v_{s}-v_{s}D_{j}\bar{\pi}_{s}, D_{i}\phi)\,ds $$ \begin{equation} \label{10.24.4} +\int_{0}^{t}I_{s\leq\tau_{m}}C_{s} \big[(\bar{\pi}_{s}v_{s},b^{i}_{s}D_{i} \phi )+ \Phi^{kr}_{s}\sigma^{ir}_{s}(\bar{\pi}_{t}D^{y}_{k}v_{s}, D_{i}\phi)\big]\,ds. \end{equation} We take a $\phi$ such that $\phi(0)=1$ and plug $\phi_{j}$ into \eqref{10.24.4} in place of $\phi$, where $\phi_{j}(x)=\phi(x/j)$, $j=1,2,...$. Observe that $$ (\zeta^{r}_{s},\phi_{j})=C_{s}(\phi_{j} ,\bar{\pi}_{s} \Phi^{kr}_{s}D^{y}_{k}v_{s}+v_{s} \Lambda^{r*}_{s}\bar{\pi}_{s}) $$ and for any $r$ and $k$ $$ \int_{0}^{T}(1,|\bar{\pi}_{s}D_{k}^{y}v_{s}|+ |v_{s}\Lambda^{r*}_{s}\bar{\pi}_{s}|)^{2}\,ds $$ $$ \leq N\int_{0}^{T}\|\bar{\pi}_{s}\|_{W^{1}_{p}}^{2} \|v_{s}\|_{W^{1}_{p'}}^{2}\,ds \leq N\|\bar{\pi}\|_{\bW^{1}_{p}(T)}^{2}<\infty, $$ where $N$ is independent of $\omega$. By the dominated convergence theorem and the rules for passing to the limit under the sign of stochastic integral it follows that in probability uniformly on $[0,T]$ $$ \int_{0}^{t}I_{s\leq\tau_{m}} (\zeta^{r}_{s},\phi_{j})\, d\tilde{w}_{s}^{r}\to \int_{0}^{t}I_{s\leq\tau_{m}} C_{s}(1,\bar{\pi}_{s} \Phi^{kr}_{s}D^{y}_{k}v_{s}+v_{s} \Lambda^{r*}_{s}\bar{\pi}_{s})\,d\tilde{w}_{s}. $$ Similarly, and in an easier fashion one analyzes the remaining terms in \eqref{10.24.4} and concludes that for any $m$ $$ d(\chi_{t},1)= C_{t}(1,\bar{\pi}_{t} \Phi^{kr}_{t}D^{y}_{k}v_{t}+v_{t} \Lambda^{r*}_{t}\bar{\pi}_{t})\,d\tilde{y}_{t},\quad t\leq \tau_{m}. $$ By using It\^o's formula we then immediately obtain that the process \eqref{11.11.6} is at least a local martingale on $[0,T]$. We rewrite it as $\xi_{t}\zeta_{t}$, where (see Remark \ref{remark 1.17.1} and Lemma \ref{lemma 11.5.1}) $ \xi_{t}= \rho_{t}(\tilde{\sfB}- \sfB_{\cdot}(0)+ \dot{\sfB}^{*}W^{-1}V,d\tilde{w}) $ and $$ \zeta_{t}= e^{-A_{t}-\int_{0}^{t}c_{s}(y_{s})\,ds} \int_{\bR^{d}}\hat{\pi}_{t}(x) v_{t}(x)\exp\big(-\tfrac{1}{2}\int_{0}^{t} |W^{1/2}_{ s}x+W^{-1/2}_{s}V_{s}|^{2}\,ds\big)\,dx. $$ Owing to Lemma \ref{lemma 10.30.1} the process $\zeta_{t}$ is bounded on $[0,T]$ by a constant times $\|\pi_{0}\|_{\cL_{p}}$ which along with Lemma \ref{lemma 10.29.2} implies that $\xi_{t}\zeta_{t}$ is a martingale. The lemma is proved. {\bf Proof of Theorem \ref{theorem 11.8.1}}. Recall that assertions (i)-(iii) are proved in Lemma \ref{lemma 10.8.2}. By Lemma \ref{lemma 10.8.02} and It\^o's formula $$ Ee^{-\int_{0}^{T}c_{s}(y_{s})\,ds}\varphi(z_{T})= Ev(0,x_{0},y_{0})=E\int_{\bR^{d}}v(0,x,y_{0}) \bar{\pi}_{0}\,dx $$ $$ =E\rho_{T}e^{-\int_{0}^{T}c_{s}(y_{s})\,ds} \int_{\bR^{d}}\varphi(x,y_{T})\bar{\pi}_{T}(x)\,dx $$ $$ =E\bar{\rho_{T}}e^{-\int_{0}^{T}c_{s}(y_{s})\,ds} \int_{\bR^{d}}\varphi(x,y_{T})\bar{\pi}_{T}(x)\,dx, $$ where $\bar{\rho_{T}}=E(\rho_{T}\mid\cF^{y}_{T})$. Since the equality between the extreme terms holds for sufficiently wide class of functions $c$, we get that $$ E\big(\varphi(z_{T})\mid\cF^{y}_{T}\big) =\bar{\rho_{T}} \int_{\bR^{d}}\varphi(x,y_{T})\bar{\pi}_{T}(x)\,dx\quad \text{(a.s.)}. $$ The arbitrariness of $\phi$ implies that $\bar{\pi}_{T}\geq0$ (a.s.) and $$ 1=\bar{\rho_{T}} \int_{\bR^{d}} \bar{\pi}_{T}(x)\,dx,\quad (1,\bar{\pi}_{T})=\int_{\bR^{d}} \bar{\pi}_{T}(x)\,dx>0,\quad \bar{\rho_{T}}=(1,\bar{\pi}_{T})^{-1} $$ (a.s.). It follows that for any Borel $f\geq0$ equation \eqref{10.13.3} holds with $t=T$. The above argument can be repeated for any $t\leq T$ by taking $t$ in place of $T$. Then we obtain \eqref{10.13.3} for any $t$. Furthermore, for any $t$ we will have that that $\bar{\pi}_{t}\geq0$ and $(1,\bar{\pi}_{t})>0$ (a.s.). Actually, the last two properties hold with probability one for all $t$ at once since with probability one $\bar{\pi}_{t}$ is a continuous $\cL_{1}$-function by Lemma \ref{lemma 10.8.2} and by Lemma \ref{lemma 10.30.1}, on the set where $\tau =\inf\{t\geq0:(1,\bar{\pi}_{t})=0\}<T$, we have $ \bar{\pi}_{T}=0$, which only happens with probability zero. The theorem is proved. {\bf Proof of Theorem \ref{theorem 11.11.1}}. We use part of notation from the proof of Lemma \ref{lemma 10.8.02} but this time take $\bar{\pi}_{t}=\eta_{t} = e^{-Q_{t}}$. Then by It\^o's formula and \eqref{9.10.5} we obtain that for each $x$ $$ d\chi_{t}=\zeta^{r}_{t}\,(d\tilde{w}^{r}_{t}+ \tilde{\sfB}^{r} \,dt)+ C_{t}(v_{t}L^{*}_{t}\bar{\pi}_{t}-\bar{\pi}_{t}L_{t}v_{t}) \,dt -C_{t}\Phi^{kr}_{t}\sigma^{ir}_{t} D_{i}(\bar{\pi}_{t}D^{y}_{k}v_{t})\,dt. $$ By using the stochastic Fubini theorem and integrating by parts we see that $$ d(\chi_{t},1)=(\zeta^{r}_{t},1)\,(d\tilde{w}^{r}_{t}+ \tilde{\sfB}^{r} \,dt) $$ which implies that process \eqref{11.11.6} is a local martingale on $[0,T]$. We rewrite it as $\xi_{t}\zeta_{t}$, where (see Remark \ref{remark 1.17.1} and Lemma \ref{lemma 11.5.1}) $ \xi_{t}= \rho_{t}(\tilde{\sfB}- \sfB_{\cdot}(0)+ \dot{\sfB}^{*}W^{-1}V,d\tilde{w}) $ and $$ \zeta_{t}= e^{-A_{t}-\int_{0}^{t}c_{s}(y_{s})\,ds} \int_{\bR^{d}} v_{t}(x)\exp\big(-\tfrac{1}{2}\int_{0}^{t} |W^{1/2}_{ s}x+W^{-1/2}_{s}V_{s}|^{2}\,ds\big)\,dx. $$ Notice that $\xi_{t}$ is a martingale and $\zeta_{t}$ is obviously bounded. By Lemma \ref{lemma 10.29.2} we conclude that process \eqref{11.11.6} is a martingale. After that it suffices to repeat the proof of Theorem \ref{theorem 11.8.1} dropping unnecessary here details concerning the fact that $(1,\bar{\pi}_{t})>0$. The theorem is proved.
\section{Introduction and statement of results} \label{sec:intro} The natural $\aut (T)$-symmetry factors through $\out (T)$, for a good number of invariants of a group $T$. The simplest examples are the abelianization, $T_{\ab}$, and the quotient of $T_{\ab}$ by its torsion, $T_{\abf}$. More well-known examples are provided by the cohomology ring $H^{\bullet} T:= H^{\bullet}(T, \C)$ and the graded Lie algebra associated to the lower central series, $\cG_{\bullet}(T):= \gr_{\bullet}T \otimes \C$. Our starting point in this work is to consider a group epimorphism, $p:\G \surj D$, with finitely generated kernel $T$, and to examine three other known types of invariants with natural outer symmetry, through the prism of the $D$-symmetry induced by the canonical homomorphism, $D\to \out (T)$. Firstly, we look at the resonance varieties $\R^i_k(T)$ (i.e., the jump loci for a certain kind of homology, associated to the ring $H^{\bullet} T$), sitting inside $H^1T$; they are reviewed in Section \ref{sec:res}, and their outer symmetry is discussed in Remark \ref{rk:dhres}. Secondly, we inspect the characteristic varieties $\V^i_k(T)$ (i.e., the jump loci for homology with rank one complex local systems), lying inside the character torus $\T (T):= \Hom (T_{\ab}, \C^*)$, and their intersection with $\T^0 (T):= \Hom (T_{\abf}, \C^*)$; their definition is recalled in Section \ref{sec:dsim}, and their outer symmetry is explained in Lemma \ref{prop:dsim}. Finally, we recollect in Section \ref{sec:rsigma} a couple of relevant facts about the Bieri-Neumann-Strebel-Renz (BNSR) invariants, $\Sigma^q(T, \Z)\subseteq H^1(T, \RR)\setminus \{ 0\}$, and we point out their outer symmetry in Lemma \ref{prop:rsim}. Our choice for the types of invariants was dictated by the fact that each of them controls a certain kind of finiteness properties. We begin with resonance, for which this relationship seems to be new. We devote Section \ref{sec:resfin} to the analysis of the connexions between triviality of resonance and finiteness of various (completed) Alexander-type invariants, for finitely generated groups. Our motivation comes from the Torelli groups, $T_g$, consisting of the isotopy classes of homeomorphisms of a closed, oriented, genus $g$ surface, inducing the identity on the first $\Z$-homology group of the surface. Their finite generation was proved by D. Johnson \cite{J1}, for $g\ge 3$. According to Hain \cite{H}, $T_g$ is a $1$-formal group in the sense of D. Sullivan \cite{Su}, for $g\ge 6$, but $T_3$ is not $1$-formal. Note also that the full mapping class group $\G_g$ is an extension of the integral symplectic group $Sp_g(\Z)$ by $T_g$, for $g\ge 1$: \begin{equation} \label{eq:deftintro} 1\to T_g \rightarrow \G_g \rightarrow Sp_g(\Z)\to 1\, . \end{equation} In particular, the resonance and characteristic varieties of $T_g$, as well as its BNSR-invariants, acquire a natural $Sp_g(\Z)$-symmetry, for $g\ge 3$. Let $K\subseteq T$ be a subgroup containing the derived group $T'$. Then $H_1K:= H_1(K, \C)$ becomes in a natural way a module over the group ring $\C T_{\ab}$, with module structure induced by $T$-conjugation. Its $I$-adic completion is denoted $\widehat{H_1K}$, where $I \subseteq \C T_{\ab}$ is the augmentation ideal. When $K=T'$, $H_1K$ is the classical Alexander invariant from link theory (over $\C$). The technique of $I$-adic completion was promoted in low-dimensional topology by Massey \cite{Mas}. Here is our first main result. \begin{thm} \label{thm:aintro} Let $T$ be a finitely generated group. \begin{enumerate} \item \label{ai1} Assume $T$ is $1$-formal. Then $\R^1_1(T) \subseteq \{ 0\}$ if and only if $\dim_{\C} \widehat{H_1T'}< \infty$. \item \label{ai2} Let $K\subseteq T$ be a subgroup containing $T'$. Then $\dim_{\C} \widehat{H_1K}< \infty$, if $\R^1_1(T) \subseteq \{ 0\}$. \end{enumerate} \end{thm} Aiming at finer finiteness properties, we go on by examining characteristic varieties, in Section \ref{sec:dsim}. Here, we start from a basic result of Dwyer and Fried \cite{DF}, as refined in \cite{PS-bns}. It says that the finiteness of Betti numbers, up to degree $q$, of normal subgroups with free abelian quotient is detected precisely by the finiteness of the intersection between characteristic varieties of type $\V^i_1$, for $i\le q$, and the corresponding connected subtorus of $\T^0$. Given the $D$-symmetry of characteristic varieties, we are thus led to consider the following context, inspired by Torelli groups. Let $L$ be a $D$-module which is finitely generated and free as an abelian group. Assume that $D$ is an arithmetic subgroup of a simple $\C$-linear algebraic group $S$ defined over $\Q$, with $\Q-\rank (S)\ge 1$. Suppose also that the $D$-action on $L$ extends to an irreducible, rational $S$-representation in $L\otimes \C$. (Note that the above assumptions are satisfied for $g\ge 3$ by $D=Sp_g(\Z) \subseteq Sp_g(\C)=S$ and $L=(T_g)_{\abf}$, due to Johnson's pioneering results on the symplectic symmetry of Torelli groups from \cite{J0, J3}.) The $D$-representation in $L$ gives rise to a natural $D$-action on the connected affine torus $\T (L):= \Hom (L, \C^*)$. \begin{thm} \label{thm:bintro} If the $D$-module $L$ satisfies the above assumptions, then $\T (L)$ is geometrically $D$-irreducible, that is, the only $D$-invariant, Zariski closed subsets of $\T (L)$ are either equal to $\T (L)$, or finite. \end{thm} We deduce Theorem \ref{thm:bintro} from a deep result in diophantine geometry, due to M. Laurent \cite{Lau}, in Section \ref{sec:dsim}. Note that the conclusion of our theorem above is in marked contrast with the behavior of the induced $D$-representation in the affine space $L\otimes \C$, for which $S$-invariant, infinite and proper Zariski closed subsets may well exist. Theorem \ref{thm:bintro} enables us to obtain in Section \ref{sec:dsim} the following consequences of the triviality of resonance. \begin{thm} \label{thm:cintro} Let $p:\G \surj D$ be a group epimorphism with finitely generated kernel $T$, having the property that $\R^1_1(T)\subseteq \{ 0\}$. Assume $D\subseteq S$ is arithmetic, where the $\C$-linear algebraic group $S$ is defined over $\Q$, simple, with $\Q-\rank (S)\ge 1$. Suppose moreover that the canonical $D$-representation in $T_{\abf}$ extends to an irreducible, rational $S$-representation in $T_{\abf}\otimes \C$. The following hold. \begin{enumerate} \item \label{vt} The intersection $\V^1_k (T) \cap \T^0 (T)$ is finite, for all $k\ge 1$. \item \label{at} If moreover $b_1(T)>1$, the Alexander polynomial $\Delta^T$ is a non-zero constant, modulo the units of the group ring $\Z T_{\abf}$. \item \label{it} For any subgroup $N\subseteq T$, containing the kernel of the canonical map, $T\surj T_{\abf}$, the first Betti number $b_1(N)$ is finite. \end{enumerate} \end{thm} Note that the computation of characteristic varieties and Alexander polynomials can be a very difficult task, in general. What makes life easier in Theorem \ref{thm:cintro}, Parts \eqref{vt} and \eqref{at}, is the arithmetic symmetry. As far as Torelli groups are concerned, we obtain the following results. Denote by $K_g$ the kernel of the canonical map, $T_g\surj (T_g)_{\abf}$, identified by Johnson in geometric terms in \cite{J2}. \begin{thm} \label{thm:dintro} Let $T_g$ be the Torelli group in genus $g\ge 4$. The following hold. \begin{enumerate} \item \label{mi1} The resonance (in degree $1$ and depth $1$) is trivial: $\R^1_1(T_g)=\{ 0\}$. \item \label{mi2} The (reduced) characteristic varieties $\V^1_k (T_g) \cap \T^0 (T_g)$ are finite, for all $k\ge 1$. \item \label{mi3} The Alexander polynomial $\Delta^{T_g}$ is a non-zero constant, modulo units. \item \label{mi4} For any subgroup $N\subseteq T_g$ containing the Johnson kernel $K_g$, the vector space $H_1(N, \C)$ is finite-dimensional. \item \label{mi5} For any subgroup $K\subseteq T_g$ containing $T_g'$, $\dim_{\C} \widehat{H_1K}< \infty$. \end{enumerate} \end{thm} We compute resonance varieties of Torelli groups in Section \ref{sec:res}. Note that $\R^1_1(T_3)=H^1 T_3$ is non-trivial; this is an easy consequence of a basic result of Hain \cite{H}, who computed the graded Lie algebra $\cG_{\bullet}(T_g)$, truncated up to degree $\bullet =2$, for $g\ge 3$. His result, formulated in the category of $Sp_g(\C)$-representations, is recorded in Section \ref{sec:csim}. We obtain the triviality of $\R^1_1(T_g)$, for $g\ge 4$, by exploiting the $Sp_g(\C)$-symmetry. Parts \eqref{mi2}-\eqref{mi3} of Theorem \ref{thm:dintro} follow from Theorem \ref{thm:cintro}. Parts \eqref{mi4}-\eqref{mi5} of Theorem \ref{thm:dintro} appear to be rather surprising, since the general belief in the literature seems to be that Torelli groups and other related groups should exhibit {\em infiniteness} properties; see for instance \cite[\S 5.1]{F} and the survey \cite{H-sur}. Theorem \ref{thm:dintro}\eqref{mi4} is a consequence of Theorem \ref{thm:cintro}\eqref{it}, and Theorem \ref{thm:dintro}\eqref{mi5} follows from Theorem \ref{thm:aintro}\eqref{ai2}. Note that the finite-dimensionality of the (uncompleted) Alexander invariant $H_1(T_g', \C)$ cannot be deduced from Theorem \ref{thm:dintro}\eqref{mi2}, since $(T_g)_{\ab}$ contains non-trivial $2$-torsion, according to Johnson \cite{J3}. To the best of our knowledge, the finiteness of $b_1(T_g')$ is an open question. Theorem \ref{thm:dintro}\eqref{mi4} solves two of the most popular finiteness problems related to Torelli groups, as recorded in Farb's list from \cite{F}: for $N=K_g$, we answer Question 5.2 (over $\C$); the general case ($K_g\subseteq N\subseteq T_g$) settles Problem 5.3 (at the level of first Betti numbers). According to \cite[erratum]{BF}, the finite generation of $K_g$ is an open question, for $g\ge 3$. Another important problem concerns the Kahler property: is $T_g$ a Kahler group (that is, the fundamental group of a compact Kahler manifold)? The answer is negative, in low genus: $T_2$ is not finitely generated, as proved by McCullough and Miller \cite{MM}, and $T_3$ violates the $1$-formality property of Kahler groups, established by Deligne, Griffiths, Morgan and Sullivan in \cite{DGMS}. Theorem 4.9 from \cite{F} states that $T_g$ is not a Kahler group for $g\ge 2$. Since the proof assumed infinite generation of $K_g$, the Kahler group problem is open as well. We investigate in Section \ref{sec:rsigma} the BNSR invariants, which control the finiteness properties of normal subgroups with abelian quotient \cite{BNS, BR}. We use their arithmetic symmetry, together with Delzant's description from \cite{D} of the first BNSR invariant of Kahler groups, and our Theorem \ref{thm:dintro}\eqref{mi1}, to prove the next result. It connects the Kahler problem for Torelli groups with the finite generation question for Johnson kernels. See Remark \ref{rem:farb} for more details on our strategy of proof. \begin{thm} \label{thm:kintro} If $T_g$ is a Kahler group, the Johnson kernel $K_g$ must be finitely generated, for $g\ge 4$. \end{thm} \section{Associated graded Lie algebra} \label{sec:csim} The symplectic symmetry is well-known to be an important tool for the study of Torelli groups. In this section, we review a basic result from Hain's work \cite{H}, related to the symplectic symmetry at the level of the associated graded Lie algebra. Let $\Sigma_g$ be a closed, oriented, genus $g\ge 1$ surface. Denote by $\G_g$ the associated {\em mapping class group}, i.e., the group of isotopy classes of orientation-preserving homeomorphisms of $\Sigma_g$. Let $Sp_g(\Z)$ be the group of symplectic automorphisms of $H:= H_1(\Sigma_g, \Z)$. The arithmetic group $Sp_g(\Z)$ is a Zariski dense subgroup of the semisimple algebraic group $Sp_g(\C)$; see e.g. \cite[Corollary 5.16]{R2}. The {\em Torelli group} $T_g$ is the kernel of the natural homomorphism $p$, which associates to an element of $\G_g$ the induced action on $H$. The defining exact sequence of the group $T_g$ is \begin{equation} \label{eq:deft} 1\to T_g \rightarrow \G_g \stackrel{p}{\rightarrow} Sp_g(\Z)\to 1\, . \end{equation} R. Hain's starting point is D. Johnson's pioneering work from \cite{J0, J1, J2, J3}, which we review first. In the sequel we will assume that $g\ge 3$. In this range, Johnson proved \cite{J1} that the group $T_g$ is finitely generated. For a group $G$, we denote by $G_{\abf}$ the quotient of its abelianization by the torsion subgroup. Among other things, Johnson \cite{J0, J3} gave a very convenient description of $(T_g)_{\abf}$, in the following way. Fix a symplectic basis of $H$, $\{ a_1,\dots, a_g,b_1,\dots, b_g \}$, and denote by $\omega= \sum_{i=1}^g a_i \wedge b_i \in \bigwedge^2 H$ the symplectic form. Note that the $Sp_g(\Z)$-action on $H$ canonically extends to a $Sp_g(\Z)$-action on the exterior algebra $\bigwedge^* H$, by graded algebra automorphisms. Consider the $Sp_g(\Z)$-equivariant embedding, $H\hookrightarrow \bigwedge^3 H$, given by $h\in H\mapsto h\wedge \omega \in \bigwedge^3 H$, and denote by $L$ the $Sp_g(\Z)$-module $\bigwedge^3 H/H$. Johnson's homomorphism, constructed in \cite{J0}, will be denoted by $J\colon T_g \to L$. \begin{theorem}[Johnson] \label{thm:john} The group homomorphism $J$ is $\G_g$-equivariant, with respect to the (left) conjugation action on $T_g$ induced by \eqref{eq:deft}, and the restriction of the $Sp_g(\Z)$-action on $L$ via $p$. It induces a $Sp_g(\Z)$-equivariant isomorphism, $J\colon (T_g)_{\abf} \isom L$. \end{theorem} Setting $H(\C)=H\otimes \C$ and $L(\C)=L\otimes \C$, it follows that the canonical representation of $Sp_g(\Z)$, coming from \eqref{eq:deft}, in $(T_g)_{\ab} \otimes \C \simeq L(\C)$, extends to a rational representation of $Sp_g(\C)$. This symplectic symmetry propagates to higher degrees, in the following sense. Recall that the {\em associated graded Lie algebra} (with respect to the lower central series) of a group $G$, $\gr_{\bullet} G$, is generated as a Lie algebra by $\gr_1 G= G_{\ab}$. \begin{lemma} \label{lem:ext} Given a group extension, \begin{equation} \label{eq:tabs} 1\to T\rightarrow \G \rightarrow D\to 1\, , \end{equation} assume that $T$ is finitely generated, $D$ is a Zariski dense subgroup of a complex linear algebraic group $S$, and the $D$-action on $T_{\ab}$ extends to a rational representation of $S$ in $T_{\ab}\otimes \C$. Then the $D$-action on $\gr_{\bullet} T$ extends to an action of $S$ on $\gr_{\bullet} T \otimes \C$, in the category of graded rational representations; moreover, every $s\in S$ acts on $\gr_{\bullet} T\otimes \C$ by a graded Lie algebra automorphism. \end{lemma} \begin{proof} Presumably this result is well-known to the experts. Being unable to find a reference, we decided to include a proof. Set $\TT_{\bullet} := \gr_{\bullet} T \otimes \C$, noting that $D$ acts on $\TT_{\bullet}$ by graded Lie algebra automorphisms. For each $q\ge 1$, denote by $K_q \subseteq \TT_1^{\otimes q}$ the kernel of the linear surjection sending $t_1\otimes \cdots \otimes t_q$ to $\ad_{t_1}\circ \cdots \ad_{t_{q-1}}(t_q) \in \TT_q$. Since $D\subseteq S$ is Zariski dense, the linear subspace $K_q$ is $S$-invariant. It remains to show that the rational representations of $S$ in $\TT_{\bullet}$ constructed in this way, which extend the $D$-action coming from \eqref{eq:tabs}, have the property that $s[a,b]=[sa, sb]$, for $s\in S$, $a\in \TT_q$ and $b\in \TT_r$. This in turn is easily proved by induction on $q$. Induction starts with $q=1$, by noting that the iterated Lie bracket, $\TT_1^{\otimes r+1}\surj \TT_{r+1}$, is $S$-equivariant by construction. For the inductive step, write $a=[t,a']$, with $t\in \TT_1$ and $a'\in \TT_{q-1}$, and use the Jacobi identity to conclude. \end{proof} By Lemma \ref{lem:ext} and Theorem \ref{thm:john}, we have a short exact sequence of rational $Sp_g(\C)$-representations, \begin{equation} \label{eq:grlow} 0\to \mathcal{K}\to \bigwedge^2 \gr_1T_g \otimes \C \stackrel{\beta}{\rightarrow} \gr_2T_g \otimes \C \to 0\, , \end{equation} where $\beta$ denotes the Lie bracket. Hain \cite{H} computed the associated exact sequence of $\sp_g$-modules, where $\sp_g$ is the Lie algebra of $Sp_g(\C)$. To describe his result, we follow the conventions from \cite[Section 6]{H}. Our references for algebraic groups (respectively Lie algebras) are \cite{HG} (respectively \cite{HL}). The Lie algebra of the maximal diagonal torus in $Sp_g(\C)$ is denoted $\h$, and has coordinates $t=(t_1,\dots, t_g)$. Let $\Phi\subseteq \h^*$ be the corresponding root system, with the standard choice of positive roots, $\Phi^+ = \{ t_i-t_j, t_i+t_j \mid 1\le i<j \le g \} \cup \{ 2t_i \mid 1\le i\le g \}$. Let $\sp_g := \s= \n^- \oplus \h \oplus \n^+$ be the canonical decomposition of the Lie algebra. We denote by $B$ the associated Borel subgroup of $Sp_g(\C):= S$, with unipotent radical $U$; the Lie algebra of $B$ is $\h \oplus \n^+$, and the Lie algebra of $U$ is $\n^+$. We work with the finite-dimensional $\s$-modules associated to rational representations of $S$. The irreducible ones are of the form $V(\lambda)$, where the dominant weight $\lambda$ is a positive integral linear combination of the fundamental weights, $\{ \lambda_j(t)= t_1+\cdots +t_j \mid 1\le j\le g \}$. It follows from Theorem \ref{thm:john} that $\gr_1 T_g \otimes \C =V(\lambda_3)$, as $\sp_g$-modules. According to \cite[Lemma 10.2]{H}, all irreducible submodules of $\bigwedge^2 V(\lambda_3)$ occur with multiplicity one, and $\bigwedge^2 V(\lambda_3)$ contains $V(2\lambda_2) \oplus V(0)$ as a submodule. \begin{theorem}[Hain] \label{thm:hainhol} The $\sp_g$-map $\beta$ from \eqref{eq:grlow} is the canonical $\sp_g$-equivariant projection of $\bigwedge^2 V(\lambda_3)$ onto the submodule $V(2\lambda_2) \oplus V(0)$. \end{theorem} \section{Resonance varieties of Torelli groups} \label{sec:res} In this section, we use representation theory to compute the resonance varieties (in degree $1$ and for depth $1$) of the Torelli groups $T_g$, for $g\ge 3$, over $\C$. We begin by reviewing the {\em resonance varieties}, $\R^i_d (A^{\bullet})$, associated to a connected, graded-commutative $\C$-algebra $A^{\bullet}$, for (degree) $i\ge 0$ and (depth) $d\ge 1$. Given $a\in A^1$, denote by $\mu_a$ left-multiplication by $a$ in $A^{\bullet}$, noting that $\mu_a^2 =0$, due to graded-commutativity. Set \begin{equation} \label{eq:defres} \R^i_d (A^{\bullet})= \{ a\in A^1 \mid \dim_{\C} H^i(A^{\bullet}, \mu_a) \ge d \}\, . \end{equation} \begin{remark} \label{rk:dhres} It seems worth pointing out that, given an arbitrary group extension \eqref{eq:tabs}, one has the following symmetry property, related to resonance. By standard homological algebra (see e.g. \cite{B}), the conjugation action of $\G$ on $T$ induces a (right) action of $D$ on $A^{\bullet}= H^{\bullet}(T, \C)$, by graded algebra automorphisms. We conclude from \eqref{eq:defres} that the $D$-representation on $A^1$ leaves $\R^i_d (A^{\bullet})$ invariant, for all $i$ and $d$. \end{remark} In this paper, we need to consider only the resonance varieties $\R_d (A^{\bullet}):= \R^1_d(A^{\bullet})$, which plainly depend only on the co-restriction of the multiplication map, $\cup :A^1\wedge A^1 \to A^2$, to its image. When $\dim_{\C}A^1 <\infty$, it is easy to see that each $\R_d (A^{\bullet})$ is a Zariski closed, homogeneous subvariety of the affine space $A^1$. For a connected space $M$, we denote $\R^i_d (H^{\bullet}(M, \C))$ by $\R^i_d (M)$. For a group $G$, $\R^i_d (G)$ means $\R^i_d (K(G, 1))$. By considering a classifying map, it is immediate to check that $\R_d(M)=\R_d(\pi_1(M))$. We will abbreviate $\R_1$ by $\R$. There is a well-known, useful relation between cup-product in low degrees and group commutator, see Sullivan \cite{S}, and also Lambe \cite{L} for details. For a group $G$ with finite first Betti number, there is a short exact sequence, \begin{equation} \label{eq:grcup} 0\to (\gr_2 G\otimes \C)^* \stackrel{d}{\rightarrow} H^1G\wedge H^1G \stackrel{\cup}{\rightarrow} H^2G \, , \end{equation} where cohomology is taken with $\C$-coefficients and $d$ is dual to the Lie bracket $\beta$ from \eqref{eq:grlow}. (Note that the only finiteness assumption needed in the proof from \cite{L} is $b_1(G)<\infty$.) In what follows, we retain the notation from Section \ref{sec:csim}. Set $V=L(\C)^*$ and $\R=\R(T_g)$. We infer from \eqref{eq:grlow} and \eqref{eq:grcup} that $\R\subseteq V$ is $Sp_g(\C)$-invariant. Here is our first step in exploiting the complex symmetry from Theorem \ref{thm:hainhol}. \begin{lemma} \label{lem:max} If $\R\ne 0$, then $\R$ contains a maximal vector of the $\sp_g$-module $V$. \end{lemma} \begin{proof} This is an easy consequence of Borel's fixed point theorem \cite[Theorem 21.2]{HG}, which guarantees the existence of a $B$-invariant line, $\C\cdot v \subseteq \R$. Invariance under the action of the maximal torus implies that $v$ belongs to a weight space of the $\h$-action on $V$, that is, $v\in V_{\lambda}$ for some $\lambda \in \h^*$. Finally, $\n^+\cdot v=0$ follows from $U$-invariance. \end{proof} Since $-1$ belongs to the Weyl group of $\sp_g$, all finite-dimensional representations of $\sp_g$ are self-dual \cite[Exercise 6 on p.116]{HL}. This remark leads to the following dual reformulation of Theorem \ref{thm:hainhol}, via \eqref{eq:grcup}. \begin{lemma} \label{lem:dual} Set $V=V(\lambda_3)$. Then $H^1T_g=V$, as $\sp_g$-modules, and the kernel of the cup-product, $\cup: \bigwedge^2 H^1T_g \to H^2T_g$, is $V(2\lambda_2) \oplus V(0)$. \end{lemma} Our main result in this section, which proves in particular Theorem \ref{thm:dintro}\eqref{mi1}, is the following. \begin{theorem} \label{thm:rest} For $g=3$, $\R(T_3)=H^1T_3$, while $\R(T_g)=0$, for $g\ge 4$. \end{theorem} The assertion for $g=3$ is an immediate consequence of Lemma \ref{lem:dual}, since in this case $\bigwedge^2V= V(2\lambda_2) \oplus V(0)$, cf. Lemma 10.2 from \cite{H}. So, we will assume in the sequel that $g\ge 4$ and $\R(T_g)\ne 0$, and use Lemma \ref{lem:max} to derive a contradiction. Set $V(0)=\C \cdot z_0$. We will need explicit maximal vectors, $v_0$ for $V(\lambda_3)$ and $u_0$ for $V(2\lambda_2)$. To this end, we recall that $V(\lambda_3)= (\bigwedge^3 H/H)\otimes \C$, where the $\sp_g$-action on $\bigwedge^{\bullet} H$ is by algebra derivations; in particular, $s\cdot x\wedge y\wedge z= s\cdot x\wedge y \wedge z+ x\wedge s\cdot y\wedge z+ x\wedge y\wedge s\cdot z$, for $s\in \s$ and $x,y,z\in H$. Set $v_0'= a_1\wedge a_2\wedge a_3$, and $u_0'= \sum_{k=3}^g ( a_1\wedge a_2\wedge a_k)\bigwedge ( a_1\wedge a_2\wedge b_k)$. Denote by $v_0$ the class of $v_0'$ in $V=V(\lambda_3)$, and let $u_0$ be the class of $u_0'$ in $\bigwedge^2V$. Using the explicit description of $\h \oplus \n^+$ from \cite[Section 6]{H}, it is straightforward to check that $v_0$ has weight $\lambda_3$, $u_0$ has weight $2\lambda_2$, and $\n^+\cdot v_0=\n^+\cdot u_0=0$. To verify that both $v_0$ and $u_0$ are non-zero, we will resort to the $Sp_g(\Z)$-equivariant contraction constructed by Johnson in \cite[p.235]{J0}, $C: \bigwedge^3 H \to H$, given by \begin{equation} \label{eq:contr} C(x\wedge y\wedge z)= (x\cdot y)z+ (y\cdot z)x+ (z\cdot x)y \, , \end{equation} where the dot designates the intersection form on $H$. Set $L'=\ker(C)$, and $L'(\C)=L' \otimes \C$. It follows from \cite[pp.238-239]{J0} that there is an induced $\sp_g$-isomorphism, \begin{equation} \label{eq:lprim} L'(\C)\isom L(\C)\, . \end{equation} Clearly, $0\ne v_0'\in L'$ and $0\ne u_0'\in \bigwedge^2 L'$. Hence, both $v_0$ and $u_0$ are non-zero, as needed. We infer from uniqueness of maximal vectors \cite[Corollary 20.2]{HL} that necessarily $v_0\in \R=\R(T_g)$, if $\R\ne 0$; see Lemma \ref{lem:max}. Denote by $W$ the $\s$-submodule $V(2\lambda_2) \oplus V(0)$ of $\bigwedge^2V$ from Lemma \ref{lem:dual}. By definition \eqref{eq:defres} and Lemma \ref{lem:dual}, $v_0\in \R$ if and only if $\im (\mu_0)\cap W \ne 0$, where $\mu_0: V\to \bigwedge^2V$ denotes left-multiplication by $v_0\in V$ in the exterior algebra. Since $\n^+\cdot v_0=0$, we conclude that $\mu_0$ is $\n^+$-equivariant. Since $v_0\in V_{\lambda_3}$, it follows that \begin{equation} \label{eq:wshift} \mu_0 (V_{\lambda'})\subseteq (\bigwedge^2V)_{\lambda_3 +\lambda'}\, , \end{equation} for each weight subspace $V_{\lambda'}\subseteq V$. \begin{lemma} \label{lem:eng} If $\R(T_g)\ne 0$, then $\im (\mu_0)\cap W$ contains a non-zero vector killed by $\n^+$. \end{lemma} \begin{proof} Taking into account the preceding discussion, the assertion will follow from Engel's theorem \cite[Theorem 3.3]{HL}, applied to the Lie algebra $\n^+$ and the $\n^+$-module $\im (\mu_0)\cap W$. It is enough to check that the action of any $n\in \n^+$ on $W$ is nilpotent. This in turn follows from the fact that $s_{\alpha_1}\cdots s_{\alpha_r}\cdot W_{\lambda'} \subseteq W_{\lambda''}$, where $\lambda''=\lambda' + \sum_{i=1}^r \alpha_i$, for any $\alpha_1,\dots ,\alpha_r \in \Phi^+$ and any non-trivial weight space $W_{\lambda'}$. Indeed, $ W_{\lambda''}=0$ for $r$ big enough. To check the vanishing claim, one may invoke \cite[Theorem 20.2(b)]{HL}, and then use a height argument. \end{proof} \begin{lemma} \label{lem:non} The following hold. \begin{enumerate} \item \label{n1} For $g\ge 4$, the vector $v_0\wedge u_0\in \bigwedge^3V$ is non-zero. \item \label{n2} The class of $( a_1\wedge a_2\wedge a_3)\bigwedge ( b_1\wedge b_2\wedge b_3)$ in $\bigwedge^2V$ does not belong to $V(0)$. \end{enumerate} \end{lemma} \begin{proof} \eqref{n1} By \eqref{eq:lprim}, it is enough to show that $v_0'\wedge u_0'\in \bigwedge^3(\wedge^3 H)$ is non-zero. This is clear, since $v_0'\wedge u_0'=\sum_{k=4}^g ( a_1\wedge a_2\wedge a_3)\bigwedge ( a_1\wedge a_2\wedge a_k) \bigwedge ( a_1\wedge a_2\wedge b_k)$. \eqref{n2} Again by \eqref{eq:lprim}, it suffices to verify that the element $e=( a_1\wedge a_2\wedge a_3)\bigwedge ( b_1\wedge b_2\wedge b_3)\in \bigwedge^2(\wedge^3 H)\otimes \C$ is not killed by $\sp_g$, since $b_1\wedge b_2\wedge b_3 \in L'$. Indeed, $T_1\cdot e= ( a_1\wedge a_2\wedge a_3)\bigwedge ( a_1\wedge b_2\wedge b_3) \ne 0$, where $T_1\in \sp_g$ is described in \cite[Section 6]{H}. \end{proof} We will finish the proof of Theorem \ref{thm:rest}, by showing that Lemmas \ref{lem:eng} and \ref{lem:non} lead to a contradiction, assuming $g\ge 4$ and $\R(T_g)\ne 0$. Indeed, in this case there is $v\in V(\lambda_3)$ such that \begin{equation} \label{eq:contra} 0\ne v_0\wedge v= w_0 + w\, , \end{equation} with $w_0\in V(0)$, $w\in V(2\lambda_2)$ and $\n^+\cdot w=0$, by Lemma \ref{lem:eng}. If $w\ne 0$, then $w\in \C^*\cdot u_0$, by uniqueness of maximal vectors. Taking the weight decomposition of $v$, it follows from \eqref{eq:wshift} that we may suppose that $w_0=0$ in \eqref{eq:contra}. Therefore, $v_0\wedge u_0=0\in \bigwedge^3V$, contradicting Lemma \ref{lem:non}\eqref{n1}. If $w=0$, then $v_0\wedge v\in \C^*\cdot z_0$, for some $v\in V(\lambda_3)_{-\lambda_3}$, by the same argument on weight decomposition as before. Since the weights $\lambda_3$ and $-\lambda_3$ are conjugate under the action of the Weyl group, the weight space $V(\lambda_3)_{-\lambda_3}$ is one-dimensional, generated by the class $\overline{v}_0$ of $b_1\wedge b_2\wedge b_3 \in L'$ in $V$; see \cite[Theorems 20.2(c) and 21.2]{HL}, \cite[Section 6]{H} and \eqref{eq:lprim}. Therefore, $v_0\wedge \overline{v}_0\in V(0)$, contradicting Lemma \ref{lem:non}\eqref{n2}. The proof of Theorem \ref{thm:rest} is thus completed. \section{Resonance and finiteness properties} \label{sec:resfin} We devote this section to a general discussion of finiteness properties related to resonance, with an application to Torelli groups. Unless otherwise mentioned, we work with $\C$-coefficients. For a graded object $\OO_{\bullet}$, the notation $\OO$ means that we forget the grading. We need to review a couple of key notions. We begin with the {\em holonomy Lie algebra} associated to a connected CW-complex $M$ with finite $1$-skeleton, $\cH_{\bullet}(M)$. This is a quadratic graded Lie algebra, defined as the quotient of the free Lie algebra generated by $H_1M$, $\bL_{\bullet}(H_1M)$, graded by bracket length, modulo the ideal generated by the image of the comultiplication map, $\partial_M: H_2M\to \bigwedge^2 H_1M$; here we are using the standard identification $\bigwedge^2 H_1M \equiv \bL_{2}(H_1M)$, given by the Lie bracket. Note that the dual of $\partial_M$ is the cup-product, $\cup_M: \bigwedge^2 H^1M\to H^2M$. When $G$ is a finitely generated group and $M=K(G,1)$, $\cH_{\bullet}(M)$ is denoted $\cH_{\bullet}(G)$. By considering a classifying map, it is easy to see that $\cH_{\bullet}(M)= \cH_{\bullet}(\pi_1(M))$, for a connected CW-complex $M$ with finite $1$-skeleton. The associated graded Lie algebra $\gr_{\bullet}G \otimes \C$, denoted $\cG_{\bullet}(G)$, is also finitely generated in degree one, but not necessarily quadratic. The canonical identification, $H_1G \equiv \cG_1(G)$, extends to a graded Lie algebra epimorphism, $\bL_{\bullet}(H_1G) \surj \cG_{\bullet}(G)$. By \eqref{eq:grcup}, this factors to a graded Lie algebra surjection, \begin{equation} \label{eq:grhol} \cH_{\bullet}(G) \surj \cG_{\bullet}(G) \, . \end{equation} Let $P_{\bullet}$ be the polynomial algebra $\Sym (H_1G)$, endowed with the usual grading. For a Lie algebra $\cH$, denote by $\cH':= [\cH, \cH]$ the derived Lie algebra, and by $\cH'':= [\cH', \cH']$ the second derived Lie algebra. The exact sequence of graded Lie algebras \begin{equation} \label{eq:binf} 0\to \cH'_{\bullet}(G)/\cH''_{\bullet}(G) \rightarrow \cH_{\bullet}(G)/\cH''_{\bullet}(G) \rightarrow \cH_{\bullet}(G)/\cH'_{\bullet}(G)\to 0 \end{equation} yields a positively graded $P_{\bullet}$-module structure on the {\em infinitesimal Alexander invariant}, $\bb_{\bullet}(G):= \cH'_{\bullet}(G)/\cH''_{\bullet}(G)$, induced by the adjoint action. A finite $P_{\bullet}$-presentation of $\bb_{\bullet}(G)$ is described in \cite[Theorem 6.2]{PS}. Let $E_{k-1}(\bb(G))\subseteq P$ be the {\em elementary ideal} generated by the codimension $k-1$ minors of a finite $P$-presentation for $\bb (G)$. Denote by $\W_k(G)\subseteq H^1G$ the zero set of $E_{k-1}(\bb(G))$. The next lemma explains the relationship between the infinitesimal Alexander invariant and the resonance varieties in degree one. \begin{lemma} \label{lem:infares} Let $G$ be a finitely generated group. Then the equality \[ \W_k (G) \setminus \{0\}=\R_k(G)\setminus \{0\} \] holds for all $k \geq 1$. \end{lemma} \begin{proof} We know from \cite[Theorem 6.2]{PS} that $\bb (G)=\coker (\nabla)$, as $P$-modules, where \begin{equation} \label{eq:bpres} \nabla:= \delta _3 + \id \otimes \partial _G\colon P\otimes\Big(\bigwedge\nolimits^3 H_1G \oplus H_2G \Big) \to P\otimes \bigwedge\nolimits^2 H_1G \, , \end{equation} and the $P$-linear map $\delta_3$ is given by $ \delta _3 (a \wedge b \wedge c) = a\otimes b \wedge c +b \otimes c \wedge a +c \otimes a \wedge b$, for $a,b,c\in H_1G$. Pick any $z \in H^1G \setminus \{0\}$. By linear algebra, we infer that $z \in \W_k (G)$ if and only if $ \dim _{\C}\coker (\nabla (z)) \geq k$. Consider the exact cochain complex $(\bigwedge ^{\bullet} H^1G, \lambda _z)$, where $\lambda _z$ denotes left multiplication by $z$, and the dual exact chain complex, $(\bigwedge^{\bullet} H_1G, {}^{\sharp} \lambda _z)$. It is straightforward to check that the restriction of $ {}^{\sharp} \lambda _z$ to $\bigwedge ^{3} H_1G$ equals $\delta _3(z)$, where $\delta _3$ is as in \eqref{eq:bpres}. Denoting by $\delta _2(z)$ the restriction of ${} ^{\sharp} \lambda _z$ to $\bigwedge ^2 H_1G$, we obtain from exactness the following isomorphism: \[ \coker (\nabla (z)) \cong \im ( \delta _2(z))/ \im ( \delta _2(z)\circ \partial _G)\, . \] By exactness again, $ \dim _{\C}\im ( \delta _2(z))=n-1$, where $n:=b_1(G)$. Hence $z \in \W_k (G)$ if and only if $\rank ( \delta _2(z)\circ \partial _G) \leq n-1-k$. Plainly, the linear map dual to $\delta _2(z)\circ \partial _G$ is $\mu _z$. Consequently, $z \in \W_k (G)$ if and only if $\rank ( \mu_z) \leq n-1-k$, that is, if and only if $z \in \R_k(G)$; see definition \eqref{eq:defres}. \end{proof} We may spell out our first general result relating resonance and finiteness. \begin{theorem} \label{thm:resinf} Let $G$ be a finitely generated group. Then $\R_1^1(G) \subseteq \{ 0\}$ if and only if $\dim_{\C} \bb(G)< \infty$. \end{theorem} \begin{proof} Lemma \ref{lem:infares} yields in particular the equality $\R_1^1(G)= \Zero (\ann \bb(G))$, away from the origin. Let $\m \subseteq P$ be the maximal ideal of $0\in H^1G$. If $\dim_{\C} \bb(G)< \infty$, then $\m^k \subseteq \ann \bb(G)$, for some $k$, by degree inspection. Taking zero sets, we obtain that $\R_1^1(G) \subseteq \{ 0\}$. Conversely, the assumption $\R_1^1(G) \subseteq \{ 0\}$ implies $\m \subseteq \sqrt{\ann \bb(G)}$, by Hilbert's Nullstellensatz. Therefore, $\m^k \subseteq \ann \bb(G)$, for some $k$. Since $\bb(G)$ is finitely generated over $P$, we infer that $\dim_{\C} \bb(G)< \infty$. \end{proof} Theorem \ref{thm:resinf} has several interesting consequences. To describe them, we first recall a couple of notions from rational homotopy theory. A {\em Malcev Lie algebra} (in the sense of Quillen) is a Lie algebra endowed with a decreasing vector space filtration satisfying certain axioms; see \cite[Appendix A]{Q}, where Quillen associates in a functorial way a Malcev Lie algebra, $\M(G)$, to an arbitrary group $G$. Following Sullivan \cite{Su}, we say that a finitely generated group $G$ is {\em $1$-formal} if $\M(G)$ is the completion with respect to degree of a quadratic Lie algebra, as a filtered Lie algebra. In Theorem 1.1 from \cite{H}, Hain proves that the Torelli group $T_g$ is $1$-formal, for $g\ge 6$ (but $T_3$ is not $1$-formal). Fundamental groups of compact Kahler manifolds are $1$-formal, as shown by Deligne, Griffiths, Morgan and Sullivan in \cite{DGMS}. Many other interesting examples of $1$-formal groups are known; see e.g. \cite{DPS} and the references therein. Next, we review the {\em Alexander invariant}, $B(G)= G'_{\ab} \otimes \C$, associated to a finitely generated group $G$. The exact sequence of groups \begin{equation} \label{eq:bgr} 1\to G'/G'' \to G/G'' \to G/G'\to 1 \end{equation} may be used to put a finitely generated module structure on $B(G)$, induced by conjugation, over the Noetherian group ring $\C G_{\ab}$. We denote by $I\subseteq \C G_{\ab}$ the augmentation ideal, and by $\widehat{B}$ the $I$-adic completion of a $\C G_{\ab}$-module $B$. Given a graded vector space $\V_{\bullet}$, $\widehat{\V}_{\bullet}$ means completion with respect to the degree filtration: $\widehat{\V}_{\bullet}= \varprojlim_{q}\V/F_q$, where $F_q= \V_{\ge q}$. The canonical filtration of $\widehat{\V}_{\bullet}$ is $\widehat{F}_q= \ker(\pi_q)$, where $\pi_q: \widehat{\V}_{\bullet}\to \V/F_q$ is the $q$-th projection of the inverse limit. The next corollary proves Theorem \ref{thm:aintro}\eqref{ai1}. \begin{corollary} \label{cor:resb} Let $G$ be a finitely generated, $1$-formal group. Then $\R_1^1(G) \subseteq \{ 0\}$ if and only if $\dim_{\C} \widehat{B(G)}< \infty$. In particular, $\R_1^1(G) \subseteq \{ 0\}$, when $\dim_{\C} B(G)< \infty$. \end{corollary} \begin{proof} Let $\m \subseteq P$ be the maximal ideal of the origin. The $1$-formality of $G$ provides a vector space isomorphism between $\widehat{B(G)}$ and the $\m$-adic completion of $\bb(G)$, cf. Theorem 5.6 from \cite{DPS}. Since $\bb_{\bullet}(G)$ is generated in degree $0$, its $\m$-adic completion coincides with the degree completion, $\widehat{\bb_{\bullet}(G)}$. Clearly, $\bb_{\bullet}(G)$ and $\widehat{\bb_{\bullet}(G)}$ are simultaneously finite-dimensional. Hence, our first claim follows from Theorem \ref{thm:resinf}. For the second claim, simply note that the completion of a finite-dimensional vector space is again finite-dimensional. \end{proof} \begin{remark} \label{rk:anca} The finitely generated group $G$ discussed in Example 6.4 from \cite{PS-bns} satisfies $\R_1^1(G) \subseteq \{ 0\}$, yet $\dim_{\C}B(G)=\infty$. It can be shown that $G$ is $1$-formal. Consequently, $\R_1^1(G) \subseteq \{ 0\}$ is only a necessary condition for the finite-dimensionality of $B(G)$, in general. In the particular case when $G$ is nilpotent, the condition $\dim_{\C} B(G)< \infty$ is automatically satisfied, since $G'$ is finitely generated. We recover in this way from Corollary \ref{cor:resb} the resonance obstruction to $1$-formality of finitely generated nilpotent groups found by Carlson and Toledo in \cite[Lemma 2.4]{CT}. We refer the reader to Macinic \cite{M}, for similar higher-degree obstructions to the formality of a finitely generated nilpotent group. As we shall see below, the main point in Corollary \ref{cor:resb} is the fact that the finite-dimensionality of $\widehat{B(G)}$ forces $\R_1^1(G) \subseteq \{ 0\}$, when $G$ is $1$-formal. We point out that $1$-formality is needed for this implication. Indeed, let $G$ be the finitely generated nilpotent Heisenberg group with $b_1(G)=2$. As noted before, $\dim_{\C} \widehat{B(G)}< \infty$. Yet, the resonance variety $\R_1^1(G)= \C^2$ is non-trivial; see for instance \cite[Proposition 5.5]{M}. \end{remark} Let $G$ be an arbitrary finitely generated group, and $K\subseteq G$ a subgroup containing $G'$. Clearly, $K$ is normal in $G$, and $G$-conjugation makes $H_1 K$ a finitely generated module over the Noetherian group ring $\C [G/K]$. By restriction via the canonical epimorphism, $G/G' \surj G/K$, $H_1 K$ becomes a finitely generated $\C G_{\ab}$-module. We may now state our next main result from this section, which proves Theorem \ref{thm:aintro}\eqref{ai2}. \begin{theorem} \label{thm:findex} Let $G$ be a finitely generated group, and $K\subseteq G$ a subgroup containing $G'$. If $\R_1^1(G) \subseteq \{ 0\}$, the vector space $\widehat{H_1K}$ is finite-dimensional. \end{theorem} \begin{proof} We first treat the particular case $K=G'$, where we know from Theorem \ref{thm:resinf} that the vector space $ \cH_{\bullet}(G)/\cH''_{\bullet}(G)$ is finite-dimensional. The canonical graded Lie algebra surjection, $\cG_{\bullet}(G) \surj \cG_{\bullet}(G/G'')$, composed with the epimorphism \eqref{eq:grhol}, gives a graded Lie algebra surjection, $\cH_{\bullet}(G) \surj \cG_{\bullet}(G/G'')$, that factors through an epimorphism \begin{equation} \label{eq:hepi} \cH_{\bullet}(G)/ \cH''_{\bullet}(G) \surj \cG_{\bullet}(G/G'') \, . \end{equation} It follows from \eqref{eq:hepi} that $\cG_{\bullet}(G/G'')=0$, for $\bullet >>0$. On the other hand, we have an isomorphism, \begin{equation} \label{eq:mass} I^q \cdot H_1G' /I^{q+1} \cdot H_1G' \simeq \cG_{q+2}(G/G'') \, , \end{equation} for $q\ge 0$; see \cite[pp.400--401]{Mas}. We infer from \eqref{eq:mass} that the $I$-adic filtration of $H_1G'$ stabilizes, for $q>>0$. Therefore, $\dim_{\C} \widehat{H_1G'}< \infty$, as asserted. For the general case, consider the extension \begin{equation} \label{eq:kext} 1\to G' \rightarrow K \rightarrow A \to 1\, , \end{equation} where $A=K/G'$ is a finitely generated subgroup of $G_{\ab}$, and denote by $(H_1G')_A$ the co-invariants of the $A$-module $H_1G'$, noting that the canonical projection, $H_1G' \surj (H_1G')_A$, is $\C G_{\ab}$-linear. The Hochschild-Serre spectral sequence of \eqref{eq:kext} with trivial $\C$-coefficients (see e.g. \cite[p.171]{B}) provides an exact sequence of finitely generated $\C G_{\ab}$-modules, \begin{equation} \label{eq:hshort} (H_1G')_A \rightarrow H_1K \rightarrow H_1A \to 0\, . \end{equation} By standard commutative algebra (see for instance \cite[Chapter 10]{AM}), the $I$-adic completion of \eqref{eq:hshort} is again exact. Since $\dim_{\C} H_1 A< \infty$, our claim about $\widehat{H_1K}$ follows from the finite-dimensionality of $\widehat{H_1G'}$. \end{proof} The next corollary proves Theorem \ref{thm:dintro}\eqref{mi5}, thanks to Theorem \ref{thm:rest}. \begin{corollary} \label{cor:kt} Let $K$ be a subgroup of $T_g$ containing $T_g'$. The $I$-adic completion of $H_1(K, \C)$ is finite-dimensional, for $g\ge 4$, where $I\subseteq \C [(T_g)_{\ab}]$ is the augmentation ideal of the group ring. \end{corollary} Denote by $K_g\subseteq T_g$ the subgroup generated by the mapping classes of Dehn twists on simple closed curves bounding in the surface $\Sigma_g$, for $g\ge 3$. Another basic result of Johnson, proved in \cite{J2}, is that $K_g= \ker(J)$, where $J:T_g \to L$ is the Johnson homomorphism. We will improve Corollary \ref{cor:kt} in the next section, when $K$ contains $K_g$. \section{Characteristic varieties of Torelli groups} \label{sec:dsim} Guided by the interplay between arithmetic and Torelli groups, coming from \eqref{eq:deft}, we will examine now groups whose characteristic varieties possess a natural discrete symmetry. We will compute the (restricted) characteristic varieties of $T_g$, in degree $1$, when $g\ge 4$, and deduce that $b_1(K_g)<\infty$. Our setup in this section is the following. Let \begin{equation} \label{eq:aext} 1\to T \rightarrow \G \stackrel{p}{\rightarrow} D\to 1 \end{equation} be a group extension, where $T$ is finitely generated and $D$ is an arithmetic subgroup of a complex linear algebraic group $S$, defined over $\Q$, simple and with $\Q$-rank at least $1$. The motivating examples are the extensions \eqref{eq:deft}, for $g\ge 3$. Under the above assumptions on $D$, we recall that any finite index subgroup $D_1\subseteq D$ is Zariski dense in $S$, as follows from Borel's density theorem; see e.g. \cite[Corollary 5.16]{R2}. We continue by reviewing a couple of relevant facts related to character tori and characteristic varieties. Let $G$ be a finitely generated group. The {\em character torus} $\T (G)=\Hom (G_{\ab}, \C^*)$ is a linear algebraic group, with coordinate ring the group algebra $\C G_{\ab}$. The connected component of $1\in \T(G)$ is $\T^0 (G)= \T(G_{\abf})$. For the beginning, we need to assume in \eqref{eq:aext} only that $T$ is finitely generated. The natural $D$-representation in $T_{\ab}$ (respectively $T_{\abf}$) induced by conjugation canonically extends to $\Z T_{\ab}$ and $\C T_{\ab}$ (respectively to $\Z T_{\abf}$ and $\C T_{\abf}$). The corresponding left $D$-action on $\T (T)$, by algebraic group automorphisms, is denoted by $d\cdot \rho$, for $d\in D$ and $\rho \in \T (T)$, and is defined by $d\cdot \rho (u)=\rho \circ d^{-1} (u)$, for $u\in T_{\ab}$. Let now $M$ be a connected CW-complex, with finite $1$-skeleton and fundamental group $G:=\pi_1(M)$. The {\em characteristic varieties} $\V^i_k (M)$ are defined for (degree) $i\ge 0$ and (depth) $k\ge 1$ by \begin{equation} \label{eq:defchar} \V^i_k (M)= \{ \rho\in \T (G) \mid \dim_{\C} H_i(M, \C_{\rho})\ge k \}\, . \end{equation} Here $\C_{\rho}$ denotes the rank one complex local system on $M$ given by the change of rings $\Z G\to \C$, corresponding to $\rho$. When $G$ is a finitely generated group and $M=K(G, 1)$, we use the notation $\V^i_k (G)$. If the complex $M$ has finite $q$-skeleton $M^{(q)}$, for some $q\ge 1$, it is easy to check that $\V^i_k (M)$ is Zariski closed in $\T (G)$, for all $i\le q$ and $k\ge 1$. It is equally easy to see that $\V^i_k (M)$ depends only on $M^{(q+1)}$, for $i\le q$ and $k\ge 1$, when $M$ has finite $1$-skeleton and $q\ge 1$. In particular, $\V^1_k (M)= \V^1_k (\pi_1(M))$, for all $k\ge 1$. Likewise, $\R^i_k (M)$ depends only on $M^{(q+1)}$, where $q\ge 1$, in the same range as before, for any $M$ with $M^{(1)}$ finite. We will be mainly interested in the characteristic variety $\V (G):=\V^1_1 (G)$ associated to a finitely generated group $G$. Our starting point in this section is the following. \begin{lemma} \label{prop:dsim} Assume in \eqref{eq:aext} that the group $T$ is finitely generated. Then the characteristic varieties $\V^i_k (T)\subseteq \T (T_{\ab})$ are $D$-invariant subsets, for all $i\ge 0$ and $k\ge 1$. \end{lemma} \begin{proof} For $\gamma \in \G$, denote $\gamma$-conjugation by $\iota_{\gamma}: T\isom T$. For a ring $R$, with group of units $R^{\times}$, and a group homomorphism, $\chi: T\to R^{\times}$, the notation $R_{\chi}$ means the $\Z T$-module $R$ associated to the change of rings $\chi: \Z T\to R$. Note that the pair $(\iota_{\gamma}, \id_R)\colon (T, R_{\chi\circ \iota_{\gamma}}) \isom (T, R_{\chi})$, gives an isomorphism in the category of local systems; see e.g. \cite[III.8]{B}. Hence, there is an induced isomorphism, $H_*(T, R_{\chi\circ \iota_{\gamma}}) \isom H_*(T, R_{\chi})$. Our claim follows by taking $R=\C$, and inspecting definition \eqref{eq:defchar}. \end{proof} Consider $\V^1_k (T)\cap \T^0(T_{\ab})$. This leads us to look at a discrete group $D$ acting linearly on a free, finitely generated abelian group $L$, and examine the $D$-invariant, Zariski closed subsets $W$ of $\T (L)= \Hom (L, \C^*)$. Besides the trivial case $W=\T (L)$, we find a lot of $0$-dimensional examples, by taking $W$ to be the subgroup of $m$-torsion elements of $\T (L)$. This raises a natural question: is there anything else? To present a first answer, we need the following notions. A {\em translated subgroup (torus)} is a subset of the character torus $\T= \T (G)$ of the form $t\cdot \bS$, where $t\in \T$ and $\bS \subseteq \T$ is a closed (connected) algebraic subgroup. Note that the {\em direction} of the translated subgroup, $\bS$, is uniquely determined by $t\cdot \bS$. A Zariski closed subset $W\subseteq \T$ is a {\em union of translated tori} if each irreducible component of $W$ is a translated torus. A fundamental result, due to Arapura \cite{A}, is that $\V (G)$ is a union of translated tori in $\T (G)$, whenever $G$ is a {\em quasi-Kahler group}, that is, the fundamental group of the complement of a divisor with normal crossings in a compact Kahler manifold. \begin{lemma} \label{prop:dchar} Let $L$ be a $D$-module which is finitely generated and free as an abelian group. Assume that $D$ is an arithmetic subgroup of a simple $\C$-linear algebraic group $S$ defined over $\Q$, with $\Q-\rank (S)\ge 1$. Suppose also that the $D$-action on $L$ extends to an irreducible, rational $S$-representation on $L(\C):= L\otimes \C$. Let $W\subset \T (L)$ be a $D$-invariant, Zariski closed, proper subset of $\T (L)$. If $W$ is a union of translated tori, then $W$ is finite. \end{lemma} \begin{proof} We know that each irreducible component of $W$ is of the form $t\cdot \bS$, as above. We have to show that $\dim (\bS)=0$. To this end, we consider the isotropy group of $t\cdot \bS$, denoted $D_1$; it is a finite index subgroup of $D$. Note that $\dim (\bS)< \dim (\T (L))$, since $W\ne \T (L)$. The $D_1$-invariance of $t\cdot \bS$ forces the direction $\bS$ to be $D_1$-invariant as well. Therefore, the Lie algebra $T_1 \bS$ is a $D_1$-invariant linear subspace of $\Hom (L, \C)=L(\C)^*$. Since $D_1\subseteq S$ is Zariski dense, it follows that the subspace $T_1 \bS$ is actually $S$-invariant. Hence, $T_1 \bS=0$, due to $S$-irreducibility. \end{proof} To improve Lemma \ref{prop:dchar}, we need a preliminary result. \begin{lemma} \label{lem:prel} Let $L$ be a $D$-module which is finitely generated and free as an abelian group. Then the subgroup $\OO_t$ of $\T (L)$, generated by the $D$-orbit $D\cdot t$, is finitely generated, for any $t\in \T (L)$. \end{lemma} \begin{proof} Pick a $\Z$-basis $\{ e_i\}$ of $L$; identify $L$ with $\Z^n$, and $\T (L)$ with $(\C^*)^n$. For $t=(t_1, \dots, t_n)\in \T (L)$ and $w= \sum_i w_i e_i \in L$, set $t^w:= \prod_i t_i^{w_i} \in \C^*$. The elements of $D\cdot t$ are of the form $(t^{v_1}, \dots, t^{v_n})$, with $v_i= \sum_j v_{ij} e_j \in L$. Define $u_k^{ij}= \delta_{ik} e_j \in L$, for $1\le i,j,k \le n$. The equality \[ \Big( t^{v_1}, \dots, t^{v_n}\Big) = \prod_{1\le i,j\le n} \Big( t^{u_1^{ij}}, \dots, t^{u_n^{ij}}\Big)^{v_{ij}} \] gives the desired conclusion. \end{proof} Now, we are going to show that the $D$-irreducibility of $L$ is inherited by the affine torus $\T (L)$, in the sense explained below. This proves Theorem \ref{thm:bintro}. \begin{theorem} \label{thm:girr} Let $L$ be a $D$-module which is finitely generated and free as an abelian group. Assume that $D$ is an arithmetic subgroup of a simple $\C$-linear algebraic group $S$ defined over $\Q$, with $\Q-\rank (S)\ge 1$. Suppose also that the $D$-action on $L$ extends to an irreducible, rational $S$-representation on $L(\C):= L\otimes \C$. Let $W\subset \T (L)$ be a $D$-invariant, Zariski closed, proper subset of $\T (L)$. Then $W$ is finite. \end{theorem} \begin{proof} We first claim that $D\cdot t$ is finite, for every $t\in W$. This is a consequence of a conjecture of Lang in diophantine geometry, proved by Laurent in \cite{Lau}. We will apply Laurent's results to the closed subvariety $W\subseteq \T (L)$ and the subgroup $\OO_t \subseteq \T (L)$, which is finitely generated, see Lemma \ref{lem:prel}. By \cite[Th\' eor\` eme 2]{Lau}, we have an inclusion \begin{equation} \label{eq:lang} W\cap \OO_t \subseteq \bigcup \gamma \cdot \bS\, , \end{equation} where the translated subgroup $\gamma \cdot \bS$ has origin $\gamma\in \OO_t$, is contained in $W$, and $\gamma \cdot \bS$ is maximal with these properties. Moreover, there are finitely many such maximal translated subgroups, according to \cite[Lemme 4]{Lau}. Denote by $V\subseteq W$ the right-hand side of \eqref{eq:lang}. Since $\OO_t$ is $D$-invariant, by construction, it follows that $V$ is $D$-invariant as well. Hence, $V$ must be finite, by Lemma \ref{prop:dchar}. Since clearly $D\cdot t\subseteq W\cap \OO_t$, we obtain our claim. Supposing that $W$ is infinite, we may find a smooth point $t\in W$ whose tangent space satisfies $0\ne T_t W \ne T_t \T (L)$. Translating to the origin, we obtain a $D_t$-invariant, proper and non-trivial linear subspace of $L(\C)^*$, where $D_t$ is the isotropy group of $t$. By the first step of our proof, $D_t$ has finite index in $D$, which contradicts the $S$-irreducibility of $L(\C)$. \end{proof} We go on by establishing a relation between characteristic and resonance varieties in degree $1$, for finitely generated groups. \begin{lemma} \label{lem:tgcone} Let $G$ be a finitely generated group. \begin{enumerate} \item \label{tc1} There is a finitely presented group $\overline{G}$, together with a group epimorphism, $\phi: \overline{G}\surj G$, such that $\phi_{\ab}: \overline{G}_{\ab}\isom G_{\ab}$ is an isomorphism, inducing identifications, $\V^1_k (\overline{G})\equiv \V^1_k(G)$ and $\R^1_k (\overline{G})\equiv \R^1_k(G)$, for all $k\ge 1$. \item \label{tc2} In particular, the tangent cone at $1$ of $\V^1_k(G)$, $TC_1 \V^1_k(G)$, is contained in $\R^1_k(G)$, for all $k\ge 1$. \end{enumerate} \end{lemma} \begin{proof} Part \eqref{tc1}. Let $X$ be a classifying space for $G$, having $1$-skeleton equal to a finite wedge of circles. Denote by $\{ \Z G_{\ab} \otimes C_{\bullet} X \stackrel{D_{\bullet}}{\rightarrow}\Z G_{\ab} \otimes C_{\bullet -1} X\}$ the cellular chain complex of the universal abelian cover $X^{\ab}$, where $\{ C_{\bullet} X \stackrel{d_{\bullet}}{\rightarrow} C_{\bullet -1} X\}$ is the cellular chain complex of $X$, over $\Z$. Since the ring $\Z G_{\ab}$ is noetherian, $\im (D_2)$ is generated over $\Z G_{\ab}$ by the images of finitely many $2$-cells, say $\{ e_1, \dots, e_r \}$. Consider now the comultiplication map, $\partial_{X^{(2)}}: H_2(X^{(2)}, \C) \to \wedge^2 H_1(X^{(2)}, \C)$. Pick finitely many $2$-cells, say $\{ e'_1, \dots, e'_s \}$, such that the $\partial$-images of the $d_{\bullet}$-cycles belonging to $\Z- \spn \{ e'_1, \dots, e'_s \}$ generate $\im (\partial_{X^{(2)}})$ over $\C$. Let $Y$ be the finite subcomplex of $X^{(2)}$ obtained from $X^{(1)}$ by attaching the cells $\{ e_i\}$ and $\{ e'_j \}$. Set $\overline{G}:= \pi_1(Y)$. Attach cells of dimension at least $3$ to $Y$, in order to obtain a classifying space for $\overline{G}$, denoted $\overline{X}$. Extend the inclusion $Y\hookrightarrow X^{(2)}$ to a map $\overline{X}\to X$, and consider the induced group epimorphism, $\phi: \overline{G}\surj G$. By our choice of the $e$-cells, we infer that $\phi$ induces in turn an isomorphism, $H_1(\overline{G}, R_{\chi\circ \phi})\isom H_1(G, R_{\chi})$, for any group homomorphism, $\chi:G \to R^{\times}$, when the ring $R$ is commutative. In particular, our claims on $\phi_{\ab}$ and characteristic varieties are thus verified. To check the claim on resonance varieties, we may replace $\overline{G}$ by $Y$ and $G$ by $X^{(2)}$. From our choice for the $e'$-cells, we deduce that, upon identifying $\wedge^2 H_1(Y, \C)$ and $\wedge^2 H_1(X^{(2)}, \C)$ via $\phi$, we have the equality $\im (\partial_{Y})= \im (\partial_{X^{(2)}})$. By duality, the cup-product maps $\cup_Y$ and $\cup_{X^{(2)}}$ have isomorphic co-restrictions to the image, which proves our last claim. Part \eqref{tc2}. Libgober proved in \cite{Lib} that $TC_1 \V^i_k(M)\subseteq \R^i_k (M)$, for $i\ge 0$ and $k\ge 1$, when $M$ is a connected finite CW-complex. This gives the desired inclusion, for $G$ finitely presented. By Part \eqref{tc1}, the inclusion still holds for finitely generated groups. \end{proof} Here is our main result in this section, relating arithmetic symmetry and finiteness properties, which proves Theorem \ref{thm:cintro}. \begin{corollary} \label{cor:qkchar} Assume in \eqref{eq:aext} that $T$ is finitely generated and $\R (T) \subseteq \{ 0\}$, and $D\subseteq S$ is arithmetic, where the $\C$-linear algebraic group $S$ is defined over $\Q$, simple, with $\Q-\rank (S)\ge 1$. Suppose moreover that the canonical $D$-representation in $T_{\abf}$ extends to an irreducible, rational $S$-representation in $T_{\abf}\otimes \C$. Then the following hold. \begin{enumerate} \item \label{tv} The restricted characteristic varieties $\V^1_k (T)\cap \T^0 (T)$ are finite, for all $k\ge 1$. \item \label{ti} The first Betti number of $\pi^{-1}(A)$ is finite, for any subgroup $A\subseteq T_{\abf}$, where $\pi: T\surj T_{\abf}$ is the canonical projection. \item \label{ta} If moreover $b_1(T)>1$, the Alexander polynomial $\Delta^T$ is a non-zero constant $c\in \Z$, modulo the units of $\Z T_{\abf}$. \end{enumerate} \end{corollary} \begin{proof} We may clearly assume $b_1(T)\ne 0$. We want to use Theorem \ref{thm:girr}, applied to the $D$-module $L=T_{\abf}$ and the closed subvariety $W=\V^1_k (T)\cap \T^0 (T)$. The $D$-invariance of $W$ follows from Lemma \ref{prop:dsim}. The fact that $W\ne \T (L)$ is a consequence of our assumption on $\R (T)$, due to Lemma \ref{lem:tgcone}\eqref{tc2}. By resorting to Theorem \ref{thm:girr}, we conclude the proof of Part \eqref{tv}. Set $K:= \ker (\pi)$. A basic result of Dwyer and Fried \cite{DF}, as refined in \cite[Corollary 6.2]{PS-bns}, says that the finiteness of $\V_1^1 (T) \cap \T^0(T)$ is equivalent to $\dim_{\C} (K_{\ab})\otimes \C <\infty$, for any finitely generated group $T$. Consider now the extension $1\to K\to \pi^{-1}(A)\to A\to 1$. A standard application of the Hochschild-Serre spectral sequence shows that the first Betti number of $\pi^{-1}(A)$ is finite, since both $b_1(K)$ and $b_1(A)$ are finite. This completes the proof of Part \eqref{ti}. Part \eqref{ta} follows from Part \eqref{tv}, via Corollary 3.2 from \cite{PS-cod}. \end{proof} Corollary \ref{cor:qkchar} leads to the following consequences, for Torelli groups, which prove Theorem \ref{thm:dintro}\eqref{mi2}-\eqref{mi4}. Let $K_g\subseteq T_g$ be the Johnson kernel from Section \ref{sec:resfin}. \begin{corollary} \label{cor:tor} Assume $g\ge 4$. \begin{enumerate} \item \label{torv} The intersection $\V^1_k (T_g)\cap \T^0 (T_g)$ is finite, for all $k\ge 1$. \item \label{tork} The vector space $H_1(N, \C)$ is finite-dimensional, for any subgroup $N$ of $T_g$ containing $K_g$. \item \label{tora} The Alexander polynomial $\Delta^{T_g}$ is a non-zero constant $c\in \Z$, modulo units. \end{enumerate} \end{corollary} \begin{proof} The condition on resonance is guaranteed by Theorem \ref{thm:rest}. Since $b_1(T_g)$ is an increasing function of $g$, for $g\ge 3$, $b_1(T_g)\ge 14$, when $g\ge 3$. \end{proof} \section{Sigma-invariants and the Kahler property} \label{sec:rsigma} We close by examining groups with natural arithmetic symmetry, at the level of Bieri-Neumann-Strebel-Renz invariants. We first review briefly the definitions and the main properties; for more details and references, see e.g. \cite{PS-bns}. We start with the {\em Novikov-Sikorav completion} of a finitely generated group $G$, with respect to $\chi\in \Hom(G, \RR)$, denoted $\widehat{\Z G}_{-\chi}$. For $k\in \Z$, let $F_k$ be the abelian subgroup of $\Z G$ generated by the elements $g\in G$ with $\chi (g)\ge k$. The completion, $\widehat{\Z G}_{-\chi}$, of $\Z G$ with respect to the decreasing filtration $\{ F_k\}_{k\in \Z}$ becomes in a natural way a ring, containing the group ring $\Z G$. The {\em Sigma-invariants} $\Sigma^q(G, \Z)$ are defined for $q\ge 1$ by \begin{equation} \label{eq:defsigma} \Sigma^q(G, \Z)= \{ 0\ne \chi\in H^1(G, \RR)\mid H_{\le q}(G, \widehat{\Z G}_{-\chi})=0 \}\, . \end{equation} When the group $G$ is of type $FP_k$, the above definition coincides with the one introduced by Bieri and Renz in \cite{BR}, for $q\le k$. Note that property $FP_1$ simply means finite generation of $G$. The Sigma-invariant $\Sigma^1(G, \Z)$ of a finitely generated group $G$, denoted $\Sigma (G)$, coincides with the one defined by Bieri, Neumann and Strebel in \cite{BNS}. It turns out that $\Sigma^q(G, \Z)$ is an open (possibly empty) conical subset of $H^1(G, \RR)$, for all $q\le k$, when $G$ is of type $FP_k$. Moreover, in this case we have the following fundamental property. Given a group epimorphism, $\nu: G\surj L$, onto an abelian group, $\ker (\nu)$ is of type $FP_q$, with $q\le k$, if and only if \begin{equation} \label{eq:ftest} \nu^*(H^1(L, \RR)\setminus \{ 0\})\subseteq \Sigma^q(G, \Z)\, . \end{equation} Here is the analog of Lemma \ref{prop:dsim} for Sigma-invariants. \begin{lemma} \label{prop:rsim} Let $p:\G \surj D$ be a group epimorphism, with finitely generated kernel $T$. Then $\Sigma^q(T, \Z)$ is invariant under the canonical action of $D$ on $H^1(T, \RR)$, coming from \eqref{eq:aext}, for all $q\ge 1$. \end{lemma} \begin{proof} Novikov-Sikorav completion is functorial, in the following sense. Let $\phi:G\to K$ be a group homomorphism and $\chi\in \Hom(K, \RR)$. The induced ring homomorphism, $\phi: \Z G\to \Z K$, clearly preserves the defining filtrations of $\widehat{\Z G}_{-\chi\circ \phi}$ and $\widehat{\Z K}_{-\chi}$. Passing to completions, $\phi$ extends to a ring homomorphism, $\widehat{\phi}: \widehat{\Z G}_{-\chi\circ \phi} \to \widehat{\Z K}_{-\chi}$. The above remark may be applied to $\gamma$-conjugation, $\phi=\iota_{\gamma}: T\isom T$, for any $\gamma\in \G$, and an arbitrary additive character $\chi\in H^1(T, \RR)$. The pair $(\phi, \widehat{\phi})$ gives then an isomorphism, $(T, \widehat{\Z T}_{-\chi\circ \phi}) \isom (T, \widehat{\Z T}_{-\chi})$, in the category of local systems. Consequently, there is an induced isomorphism, $H_*(T, \widehat{\Z T}_{-\chi\circ \phi}) \isom H_*(T, \widehat{\Z T}_{-\chi})$. We infer from \eqref{eq:defsigma} that $\chi\in \Sigma^q(T, \Z)$ if and only if $\chi\cdot d\in \Sigma^q(T, \Z)$, where $d=p(\gamma)$. \end{proof} \begin{remark} \label{rem:anti} Note that $-\id \in Sp_g(\Z)$ acts by $-\id$ on $\wedge^3 H/H$. Consequently, $-\Sigma (T_g)=\Sigma (T_g)$. This symmetry property of $\Sigma (G)$ about the origin does not hold in general. Note also that, when $-\Sigma (G)=\Sigma (G)$, $\Sigma (G)\ne \emptyset$ if and only if there is a finitely generated, normal subgroup $N$ of $G$, with infinite abelian quotient $G/N$. Moreover, in this statement $G/N$ may be replaced by $\Z$. Indeed, assuming $G/N$ to be infinite abelian, we infer that $\Sigma (G)\ne \emptyset$, by resorting to \eqref{eq:ftest}. Conversely, we know that the image of $\Sigma (G)$ in the quotient sphere, $(H^1(G, \RR)\setminus \{ 0\})/\RR_{+}$, is open and nonvoid. The density of rational points on this sphere \cite[p.451]{BNS} implies then that $\Sigma (G)$ contains an epimorphism, $\nu:G \surj \Z$. By antipodal symmetry of $\Sigma (G)$ and \eqref{eq:ftest} again, $\ker (\nu)$ must be finitely generated. \end{remark} Under additional hypotheses, Lemma \ref{prop:rsim} may be used to obtain strong information on finiteness properties. \begin{prop} \label{cor:bnsa} Let $p:\G \surj D$ be a group epimorphism, with finitely generated kernel $T$. Assume that $D$ is an arithmetic subgroup of a complex, simple linear algebraic group $S$, defined over $\Q$, with $\Q-\rank (S)\ge 1$, and the canonical $D$-action on $T_{\abf}$ induced by conjugation extends to a non-trivial, irreducible rational representation of $S$ in $T_{\abf}\otimes \C$. Suppose moreover that \begin{enumerate} \item \label{r1} either $\Sigma(T)$ is a finite disjoint union of finite intersections of open half-spaces in $H^1(T, \RR)$, \item \label{r2} or $\Sigma(T)$ is the complement of a finite union of linear subspaces in $H^1(T, \RR)$. \end{enumerate} Then the kernel $K$ of the natural map, $T\surj T_{\abf}$, is finitely generated if and only if $\Sigma(T)\ne \emptyset$. \end{prop} \begin{proof} According to \eqref{eq:ftest}, finite generation of $K$ is equivalent to $\Sigma(T)= H^1(T, \RR)\setminus \{0\}$. Note that our irreducibility assumptions imply that $b_1(T)>1$. We claim that if $\Sigma(T)$ is a proper, non-void subset of $H^1(T, \RR)\setminus \{0\}$, then there is a proper, non-trivial linear subspace $E\subseteq H^1(T, \RR)$ invariant under the canonical action of a finite index subgroup $D_0\subseteq D$. Granting the claim, we may use the fact that $D_0$ is Zariski dense in $S$ to infer that $E\otimes \C\subseteq (T_{\abf}\otimes \C)^*$ is $S$-invariant, a contradiction. Thus, we only need to verify the above claim, in order to finish the proof. \eqref{r1} In this case, we know that $\Sigma(T)= \cup_{i=1}^r C_i$, where each $C_i$ is a chamber of a non-void, finite hyperplane arrangement $\A_i$ in $H^1(T, \RR)$, and the union is disjoint. By Lemma \ref{prop:rsim}, there is a finite index subgroup $D_1\subseteq D$ such that $C_1\cdot d=C_1$, for any $d\in D_1$. Consider the supporting hyperplanes of $C_1$, that is, the set $\SS_1$ consisting of those hyperplanes $E\subseteq H^1(T, \RR)$ with the property that the intersection of $E$ with the boundary of $C_1$ has non-void interior in $E$. Standard arguments show that $\SS_1$ is a non-void subset of $\A_1$; see \cite[Chapter V.1]{Bo}. Clearly, $\SS_1$ is $D_1$-invariant, so we may choose $D_0$ to be the isotropy group of an element $E\in \SS_1$. \eqref{r2} In the second case, the complement of $\Sigma(T)$ is the union of a non-void, finite arrangement $\A$ of non-trivial, proper linear subspaces of $H^1(T, \RR)$. Again by Lemma \ref{prop:rsim}, $\A$ is $D$-invariant. Clearly, the isotropy group $D_0$ of $E\in \A$ satisfies the desired conditions. \end{proof} Property \eqref{r1} above is verified by $3$-manifold groups (that is, fundamental groups of compact, connected, differentiable $3$-manifolds), according to \cite{BNS}. Property \eqref{r2} holds for Kahler groups, as shown by Delzant in \cite{D}. For Torelli groups, Proposition \ref{cor:bnsa}\eqref{r2} may be improved to obtain Theorem \ref{thm:kintro}. \begin{corollary} \label{cor:ftor} If the Torelli group $T_g$ ($g\ge 4$) is a Kahler group, then the Johnson kernel $K_g$ is finitely generated. \end{corollary} \begin{proof} We have to show that $\Sigma(T_g)\ne \emptyset$, for $g\ge 4$, assuming the Kahler property. According to \cite{D}, the complement of $\Sigma(T_g)$ is a finite union, \[ \bigcup_{\alpha} f^*_{\alpha} H^1(C_{\alpha}, \RR)\, , \] coming from irrational pencils on the compact Kahler manifold $M$, if $T_g=\pi_1(M)$. More precisely, each $f_{\alpha}:M \to C_{\alpha}$ is a holomorphic map onto a smooth compact curve with $\chi (C_{\alpha})\le 0$, having connected fibers, and $\dim_{\RR} f^*_{\alpha} H^1(C_{\alpha}, \RR)= b_1(C_{\alpha})$. If $b_1(C_{\alpha})<b_1(T_g)$, for every $\alpha$, then we are done. Otherwise, $\chi (C_{\alpha})< 0$, for some $\alpha$, since $b_1(T_g)>2$. We infer from definition \eqref{eq:defres} that $\R^1_1(C_{\alpha})=H^1(C_{\alpha}, \C)$ and $f^*_{\alpha} H^1(C_{\alpha}, \RR) \subseteq \R^1_1(T_g)\cap H^1(T_g, \RR)$, which contradicts Theorem \ref{thm:rest}. \end{proof} \begin{example} \label{rem:ell} We point out that there exist Kahler groups with $\R^1_1(G)=0$ and $\Sigma(G)= \emptyset$. In particular, \eqref{eq:ftest} implies that in this situation the kernel of the canonical projection, $G\surj G_{\abf}$, is not finitely generated. At the same time, the condition on $\R^1_1(G)$ implies that there is no group epimorphism, $G\surj \pi_1(\Sigma_h)$, when the genus $h$ is at least $2$, by the argument in the proof of Corollary \ref{cor:ftor}. We give such an example, inspired by a construction of Beauville \cite[Example 1.8]{Be}. Let $g$ be a fixed-point free involution of a $1$-connected compact Kahler manifold $E$. The existence of such an object follows for instance from Serre's result \cite{Se}, which guarantees the realizability of finite groups as fundamental groups of smooth, projective complex varieties. Let $\Z_2$ act on the Fermat curve $F:= \{ x^4+y^4+z^4=0\}\subseteq \C \PP^2$ by $g(x:y:z)=(y:x:z)$. Set $M:= F\times E/\Z_2$, where the quotient is taken with respect to the diagonal action, and $C:=F/\Z_2$. Note that $M$ is a compact Kahler manifold, and $C$ is an elliptic curve. Set $G:= \pi_1(M)$. The first projection induces a holomorphic surjection, $f:M\to C$, having connected fibers, and $4$ multiple fibers (of multiplicity $2$). By Delzant \cite{D}, the subspace $f^*H^1(C, \RR)$ is contained in the complement of $\Sigma(G)$. Since $f^{\bullet}:H^{\bullet}C \to H^{\bullet}M$ may be identified with the inclusion of fixed points, $(H^{\bullet}F)^{\Z_2} \hookrightarrow (H^{\bullet}F \otimes H^{\bullet}E)^{\Z_2}$, and $b_1(E)=0$, we infer that $f$ induces in cohomology an isomorphism in degree $1$ and a monomorphism in degree $2$. It follows that $\Sigma(G)= \emptyset$ and $\R^1_1(G)=0$, as asserted. By a similar construction, we may exhibit examples of Kahler groups with arbitrary (non-zero) even first Betti number, having the property that $\R^1_1(G)=0$ and $\Sigma(G) \ne H^1(G, \RR)\setminus \{ 0\}$. As before, these conditions imply that the kernel of the canonical projection, $G\surj G_{\abf}$, is not finitely generated, and there is no group epimorphism, $G\surj \pi_1(\Sigma_h)$, for $h\ge 2$. \end{example} \begin{remark} \label{rem:farb} In the proof of Corollary \ref{cor:ftor}, our strategy involves two steps. Firstly, group surjections $T_g\surj \pi_1(\Sigma_h)$ with $h\ge 2$ are ruled out with the aid of our result on resonance, from Theorem \ref{thm:rest}. Secondly, we use the symplectic symmetry of Sigma-invariants to prove non-existence of group surjections onto orbifold fundamental groups in genus $1$ (or, elliptic pencils with multiple fibers, in the geometric language from \cite{D}), $T_g\surj \pi_1^{\orb}(\Sigma_1)$. Example \ref{rem:ell} shows that the second step is needed for the proof of Corollary \ref{cor:ftor}. \end{remark} \begin{ack} We are grateful to Alex Suciu, for useful discussions at an early stage of this work. \end{ack} \newcommand{\arxiv}[1] {\texttt{\href{http://arxiv.org/abs/#1}{arXiv:#1}}} \bibliographystyle{amsplain}
\section{Introduction} Stars of different stellar generations\footnote{We will use the words stellar ``generation" and ``population" interchangeably, for the reasons explained later.} are currently routinely found in all Galactic globular clusters (GCs). A common misunderstanding is to identify {\it multiple sequences} in the color-magnitude diagrams (CMDs) as the {\it only} evidence of {\it multiple stellar populations} in GCs, while they are only the tip of the iceberg, observable when differences in chemical composition and/or age are very large. However, all GCs display large spreads in abundances of light elements. These are due to the presence of two generations of stars, separated by a small difference in age, not enough to appear as a smearing of the cluster turn-off, but clearly visible as a different chemical signature. This difference can be uncovered by spectroscopy (see \citealt{gratton04} for a recent review). Since its discovery by the Lick-Texas group (see the early review by \citealt{kraft94}), the Na-O anticorrelation in GCs offers a powerful instrument to resolve age differences as small as a few 10$^7$ yrs (should very massive stars be involved) by means of a signal as large as 1 full dex in abundance. {\it Whenever the Na-O anticorrelation is observed in a GC, multiple stellar generations are present in the cluster}. The primordial P component of first generation stars, present in almost all Galactic GCs surveyed so far, has the high [O/Fe] and low [Na/Fe] ratios typical of core-collapse supernovae (SNe II) nucleosynthesis, characteristic of halo field stars with similar metallicity. On the other hand, \citet{carretta09a} showed that two other components of second generation stars may be found: one dominant component has intermediate composition (I; observed in $each$ GC), and another has extremely (E) modified composition, showing large O-depletion and Na (and Al) enhancements, present only in the most massive clusters\footnote{The operational distinction between I and E components \citep{carretta09a} is not relevant here and we only consider the total fraction (I+E) of second generation stars.}. This general pattern can be qualitatively produced by proton-capture reactions in H-burning at high temperature (\citealt{dd89,langer93}). The fact that the same Na-O and Mg-Al anticorrelations are found in unevolved stars in several GCs (\citealt{gratton01,rc02,carretta04}) tells us that this composition is inherited from the ashes of a previous generation of stars. While the involved processes have been largely identified, it is still not clear {\it where} these reactions occurred, the favorite sites being either main-sequence core H-burning of fast rotating massive stars (FRMS, \citealt{decressin07}) or the hot bottom of the convective envelope in intermediate-mass asymptotic giant branch (AGB) stars \citep{ventura01}. Whichever the sites are, they can not synthesize iron or other elements heavier than Al. Indeed, apart from a few notable exceptions ($\omega$~Cen, see reference in \citealt{gratton04}; M~22, \citealt{marino09,dacosta09}; M~54, \citealt{carretta10a}; Terzan~5, \citealt{ferraro09}; and possibly NGC~1851, \citealt{yg08}) the level in [Fe/H] or $\alpha$ and Fe-group elements is very constant in GCs \citep{carretta09b}. However, a recent work by \citet[from now L09]{lee09} seems to challenge this consolidated scenario, because in their Ca-$uvby$ survey they found a spread in the photometric $hk$ index (including Ca {\sc ii} H and K lines) for several GCs. They interpret this spread as due to Ca abundance variations and claim that this $\alpha-$element was produced by SN II in a past phase of the lifetime of GCs. In turn, these should be assumed to be initially much more massive than at present, else the highly energetic SNe winds would have been lost from the potential well. Moreover, they find that the Ca-strong group in GCs with evidence of multiple stellar populations from spectroscopy is associated to the population of stars with lower O and higher Na values, and vice versa concerning the Ca-weak stars. The need of massive proto-clusters is common to most scenarios of chemical evolution of GCs, including one proposed by us \citep{carretta10b}. However a correlation between Ca and light elements would imply that the second generation stars were formed from material enriched not only by FRMS or AGB stars, but also by SN II events, which is a substantial modification of the current scenario. In the present Letter we exploit the large wealth of data from our ongoing project ``Na-O anticorrelation and HB" \citep{carretta06} and from other extensive spectroscopic surveys to compare [Ca/H] ratios obtained from high resolution spectroscopy with the abundance distribution of O, Na, Al in a sample of several hundred red giant branch (RGB) stars in 17 GCs. This allows to shed light on the r\^ole of Ca in the chemical evolution of GCs; from our extensive spectroscopic database, we seek to confirm whether the spread in the photometric $hk$ index is driven by a real spread in Ca abundance. \section{Atmospheric parameters and abundance analysis} The full description of the analysis is given in \citet{carretta09b}; here we present only the abundances of Ca, which are based on a large number of lines (typically 10-12 in the most metal-poor clusters like M~15 and almost 20 in metal-rich GCs like 47~Tuc). The [Ca/Fe] ratios with the number of lines and the line-to-line r.m.s. scatter for each star are listed in Table \ref{cah} (completely available on line only). In Table 2 we report the number of stars in each cluster, the average [Ca/Fe] and [Ca/H] values and the 1~$\sigma$ scatter of the mean. The rms scatters in [Ca/Fe] are smaller than in [Ca/H] because the sensitivity of Ca and Fe lines to atmospheric parameters is quite similar. Five of the clusters listed in Table 1 are included in the sample by L09, who claim significant spread in Ca for four of them. The values listed in Table~2 indicate that the mean level of Ca is highly homogeneous in each GCs. The average r.m.s. star-to-star scatter for [Ca/Fe] is 0.03~dex from 17 GCs, to be compared with average scatter of 0.18 dex, 0.22 dex, 0.07 dex and 0.25 dex for O, Na, Mg, and Al respectively. Note that the mean scatter in Mg, produced by core collapse SNe but also involved in proton-capture reactions, is more than {\it twice} that of Ca, which does not participate in any H-burning. \section{Results} Since we have only about 10 stars per cluster observed with UVES, we combined the 17 GCs, separating first generation (P) stars from second generation (I+E) stars. For direct comparison with L09, we used [Ca/H] instead of [Ca/Fe], after normalizing the individual values to the mean of each cluster. In Figure 1 we show the histogram in [Ca/H] for P and I+E stars, with average values, errors, and rms. About two thirds of stars belong to the second generation, and the average normalized [Ca/H] values do not differ from those of first generation (at about 1$\sigma$). A further check comes from the cumulative distributions, shown in Figure 2 (upper panel), from which we excluded the 12 stars of NGC~2808, the GC in our sample showing the most extreme variations in He (\citealt{piotto07,bra10}, and below in this Letter). The two distributions cannot be distinguished. Note that if we include also NGC~2808, the probability of the Kolmogorov-Smirnov drops to 0.07, still compatible with no significant difference, but possibly indicating some actual variations in this particular cluster. However, when we show in the three lower panels of Figure 2 the cumulative distributions for [O/Fe], [Na/Fe], and [Al/Fe], all elements involved in the (anti)correlations in GCs, we clearly see a striking difference (compare e.g. to supplementary Fig.14 of L09). Our sample of 200 stars comes from 17 different GCs, but there is already available in literature a sample of similar size for a single GC, M~4, for which \citet{marino08} analyzed high resolution UVES spectra of 105 RGB stars. This cluster is also in the L09 sample; from their supplementary Fig.14 (panel i) we see that almost all stars defined Ca-strong are Na-rich (in our interpretation, second generation I, E) and vice versa for the Ca-weak. However, when we divide the stars in the Marino et al. sample in P and I+E (using their same separation at [Na/Fe]=0.2) and compare the histograms and the means of the two populations (see Figure 3, left panel), they look very similar\footnote{As noticed by the referee, the same result holds using the 38 stars in NGC~6752 analyzed by \citet{yong05}. }. This is supported by the cumulative distributions (Figure 3, right panel): they are indistinguishable, according to the KS test. A further check was done considering only stars in the two extreme quartiles in Na abundance: the average Ca values in the more Na-poor and in the more Na-rich stars perfectly agree with the ones in Figure 3 for the I+E and P stars, respectively. {\em The stars called Ca-weak (Ca-strong) on the basis of their $hk$\ index do not seem to have lower (higher) [Ca/H] values from direct analysis of high resolution spectra; this demonstrates that the differences in the $hk$\ index must be produced by something other than Ca variations.} It is important to stress that the observational effect at the basis of the L09 claim corresponds to 1$\sigma$ spreads in [Ca/H] of about 0.03 dex in the most extreme cases (NGC~1851, NGC~2808, excluding the special cases of $\omega$ Cen and M~22 where the presence of a significant spread in iron has already been established with high resolution spectroscopy) and of about 0.02 dex for typical cases (like M~4, M~5, and NGC~6752, see Table 2)\footnote{We used their supplementary Table 3 for the FWHM of the RGB of the GC we have in common, converted them back to scatter in $hk$ assuming a Gaussian distribution, and further converted to scatter in [Ca/H] and [Fe/H] using the sensitivity of Ca on the variation in $hk$ (suppl. Sect. 4) and the relations in Sect. 3.}. In our view, these numbers call for two basic considerations: \begin{enumerate} \item Such tiny spreads are similar in amplitude to several uncertainties that are known to affect the abundance analysis (like e,g, variations in the He content); they are usually neglected since the overall uncertainty on single measures is typically larger than this. We do not exclude that there might be dispersions at these very low levels, but presently they can not be separated from the intrinsic errors. \item even if we regard these very small variations as {\em real}, some basic algebra indicates that, assuming a mass of $2\times 10^5~M_{\sun}$ and a metallicity of $Z=10^{-3}$ for the gas cloud that gave origin to the cluster, a single SN II event would be sufficient to produce the required amount of Ca (and Fe). \end{enumerate} \section{The case for NGC~1851} If the tiny variations detected in most GCs are not measuring intrinsic scatter in Ca, what might be the cause for the variations in the $hk$ index? L09 discussed several issues to show that the only way to affect the $hk$ index is a change in Ca. Specifically, they made a strong case for NGC~1851, where the few ``Ca-strong" stars (with abundances from high resolution spectroscopy by \citet{yg08} are segregated to the reddest positions in the $V,hk$ CMD. In a recent paper, \citet{han09} presented a split of the RGB in this cluster in the $U,U-I$ plane, where the redder sequence is populated by Ca-strong stars. We simply note the striking similarity between the CMD by Han et al. with the $U,U-B$ CMD for M~4 by \citet{marino08}. In both cases, the RGB is clearly split into two distinct sequences. However, the same dichotomy of RGB sequences in M~4 is explained by \citet{marino08} as the N enhancement affecting the $U-B$ colors (through the CN and NH molecular bands within the $U$ filter), so that Na-rich (and likely N rich) stars are separated from Na-Poor (and N-poor) stars in the CMDs. Since we demonstrated in Figure~3 that in M~4 the [Ca/H] ratio is identical for the Na-poor and Na-rich stars, the same conclusion might be valid for NGC~1851. A further piece of the puzzle comes from \citet{carretta09a}, who confirmed results by \citet{yong08} in NGC~6752: Na-poor stars define a narrow N-poor sequence all confined to the bluest part of the RGB in NGC~6752, whereas the Na-rich (N-enhanced) stars are spread out to the red part of the RGB, reaching large N abundances. Now, regardless of the nature of the polluters, the ejected matter is always enriched in the main outcome of H-burning, He. \citet[in preparation]{bra10} used a large sample (about 1400 RGB stars in 19 GCs) to show that stars with higher He content should have on average slightly larger [Fe/H] values the difference between the E and P stars being 0.027 dex in [Fe/H], even if there is no difference in the overall metal content Z. These findings suggest that the correlation observed by L09 between the (small) dispersion in [Ca/H] and the (large) spread in $hk$ may be related to variations in He and light elements only, with no fresh production of Fe and Ca. The sizable effect on the Kolmogorov-Smirnov test of the inclusion of NGC~2808 in the sample (Sect.~3) lends further support to this scenario, as this cluster seems to display especially large variations in He abundance \citep{piotto07,bra10}. Hence, while it is quite possible that in NGC~1851 some small intrinsic scatter in iron does actually exist \citep{yg08}, our results do not support the idea that core collapse SNe contributed to the enrichment of second generation stars in most GCs, and indicate that the production of the proton-capture elements occurred in sites not able to synthesize Ca and Fe-peak elements. \section{Discussion and conclusions} Lee and coworkers devised a scenario for chemical enrichment in clusters with multiple populations. It includes two-phases. First, core collapse SNe from first generation stars inject in the intra-cluster medium material enriched with Fe and $\alpha-$elements; at least part of this material is not lost by the GCs owing to their larger mass at early epochs. Second generation stars form from this enriched material, incorporating the ejecta of intermediate-mass AGB stars with longer evolutionary times. The second generation stars are then produced from matter enriched both in heavy and light elements. However, in our opinion, this scenario has to face several problems: \begin{itemize} \item Since there is clear evidence of multiple populations from spectroscopy in {\em all} GCs, except perhaps a few small ones with only a handful of stars analyzed (e.g. Pal 12, \citealt{cohen04}; Ter~7, \citealt{sbordone07}; both belonging to the Sagittarius dwarf galaxy), any proposed scenario must account for the formation of the generality of GCs. \item While a single core collapse SN might be enough to produce the claimed scatter in [Ca/H], thousands (or even more) FRMS or massive AGB stars are required to produce the observed pattern for elements involved in proton-capture processes. The impact of these two mechanisms can not be the same. \item Core collapse SNe should produce a (variable) enhancement in the mean abundances for other $\alpha-$elements. No significant variation has been observed for Mg and Si (apart from those explained by the Mg-Al cycle; see Fig. 3 in \citealt{carretta09b}), or Ti \citep{gratton04}. \item In GCs like M~4, a split in $hk$ of 45$\sigma_p(hk)$ (measurement errors for the populations, L09, supplementary Fig. 10) between two populations translates in virtually no difference in the [Ca/H] ratios obtained from a homogeneous large sample of stars with high resolution spectra. \item If FRMS are the polluters, the whole model would fail, since these stars release the polluted matter in very short time, while still in MS. Hence, the timescale is even shorter than that for production of core collapse SNe. \item The model by L09, adapted to explain the double RGB seen in NGC~1851, results in a number ratio of first-to second generation stars equal to 3 (75\%/25\%). From spectroscopy, on the other hand, it was found that the bulk of stars in GCs belong to the second generation \citep{cohen02,carretta09b}, with a number ratio typically of about 0.5 (0.33/0.66). We stress that this feature is $not$ a prerogative of mono-metallic GCs, the same value for the primordial fraction is also found in $\omega$ Cen and M~54 \citep{carretta10a}, the two most massive clusters in the Galaxy, both showing actual (large) dispersion in [Fe/H]: this finding demonstrates that the relative fractions of first and second generation stars are not related to the enrichment of heavy elements. \item The $r-$process element europium, whose main production site is expected to be core-collapse SNe, is fairly constant within all the studied globular clusters \citep{gratton04}. \item Finally, one of the strongest objection to the L09 model comes from the run of Al. With its yield extremely dependent on the neutron excess (metallicity), any variations in the content of Fe (or heavy element produced in SN II, like Ca) should be accompanied by clear changes in the primordial Al abundance. As shown by \citet{carretta09b} this is not always seen, in particular in clusters like M~4 where a bimodal distribution is claimed in the $hk$ distribution. The Al distribution is far from being bimodal in M~4, on the contrary [Al/Fe] is quite constant and does not vary between first and second generation stars. This is clearly seen also in other GCs like NGC~6171, NGC~288, and NGC~6838 from our own data \citep{carretta09b} as well in the larger sample by \citet{marino08}. This constancy, due to the secondary (metallicity-dependent) production of Al in SNe strongly precludes that any significant difference in heavy metal-content is provided by SNe to stars in different stellar generations in GCs. \end{itemize} We conclude that (i) the spreads in [Ca/H] are very small; (ii) different distributions in O, Na, Al are clearly seen between first and second generation stars in $all$ GCs analyzed so far, on the contrary no statistically significant difference is found regarding Ca abundances from high resolution spectroscopy; (iii) the small difference in Ca observed in some cases might be well explained with systematic second-order effects due to the fact that second generation stars are He-enhanced, and appear more metal-rich and hotter when analyzed with high resolution spectroscopy. At the moment, we cannot offer a viable alternative explanation for the dispersion in the $hk$ index observed by Lee and collaborators, although this issue certainly requires further studies. \acknowledgments Partial funding come from PRIN MIUR 2007, PRIN INAF 2007, and the DFG cluster of excellence ''Origin and Structure of the Universe''.
\section{Introduction} In their seminal paper \cite{AharonovBohm59}, Y. Aharonov and D. Bohm claim that, contrary to the conclusions of classical mechanics, the electromagnetic 4-potential $A$ affects the motion of an electron beam, even in regions where the electromagnetic field vanishes. They proposed two kinds of experiments which were successfully performed later (see, for instance, \cite{OsakabeMatsudaKawasakiEndoTonomura86} and \cite{van Oudenaarden}). In \cite{AharonovBohm59} a scalar function $\mathcal{S}$ such that $\nabla \mathcal{S}=(e\hbar/c)A$ is introduced. It has been shown that if $\psi_0$ is the solution of the Schr\"odinger equation in the absence of an EM field, then the function $\psi=\psi_0e^{-i\mathcal{S}/\hbar}$ is the solution of the equation in the presence of the field. When the region is multiply connected, $\mathcal{S}$ is not a single-valued function, and calculating $\mathcal{S}$ by two non-equivalent paths can produce $\mathcal{S}_1-\mathcal{S}_2\neq 0$. It is argued that in the magnetic AB experiment, the difference of values of some multivalued function $\mathcal{S}(x)$ is measured. The function $\mathcal{S}(x)$ is a real-valued function of spatial coordinates and is the logarithm of the phase factor of the corresponding Schr\"odinger equation. The ability of the four-potential $A$ to account for the Aharonov-Bohm effect is limited to the case in which the EM field is zero. In this paper, we give a more precise explanation of the effect by shifting the focus from the four-potential $A$ to a complex-valued multifunction $S$. Obviously, an everywhere defined \textit{real}-valued function $S$ satisfying $\nabla S=(e\hbar/c)A$ cannot produce a non-trivial field, since in this case the expression $F=\nabla\times A$ vanishes. But, as we have shown in \cite{FGW}, we can describe an electromagnetic field by a \textit{complex}-valued function $S(x)$, which we call the \textit{scalar complex potential of the electromagnetic field}. The real and imaginary parts of $S(x)$ are, in fact, the two 'scalar potentials' introduced by E.~T.~Whittaker in 1904 \cite{Whittaker04}. Note that another complex scalar potential was introduced by H.~S.~Green and E.~Wolf in 1953 \cite{GreenWolf53}. They described the similarity of the expressions for energy and momentum densities between their potential and the wave function. The relation of this potential to the Whittaker one is still unclear. Let us describe the function $S(x)$ in more details. Given a moving charge and an observer, we obtain a complex dimensionless scalar. We then define the scalar complex potential $S$ as the logarithm of this dimensionless scalar. As in \cite{AharonovBohm59}, our $S$ is not a multi-valued function (it contains complex logarithm). Moreover, we shall see that the multi-valued nature of $S$ is the \textit{tailor-made} mathematical expression of the Aharonov-Bohm effect. Next, we derive the complex Faraday vector $\mathbf{F}:=\mathbf{E}+i\mathbf{B}$ of an electromagnetic field from the pre-potential $S$: \[ F_j={\partial}^\nu (\alpha_j)_\nu ^\lambda{\partial}_\lambda S,\] where the $\alpha_j$ are Dirac's $\alpha$-matrices. These matrices $\alpha_k$ are used to insert a Lorentz invariant conjugation between the gradient and the curl as they are applied to $S$. Finally, we presnt a third-order differential equation expressing the connection between the complex potential $S(x)$ and the field sources. \section{Definition of the Complex Potential of a moving charge} Denote by $P=(t,x,y,z)$ a point in space-time at which we want to calculate the four potential. We call $P$ the observer. Denote by $L$ the world-line of the charge $q$ generating our electromagnetic field. Let $Q\in L$ be the unique point of intersection of the past light cone at $P$ with the world-line $L$ of the charge. We denote the time of the event $Q$ by $\tilde{\tau}$ and refer to this time as the \textit{retarded time} of the potential. Note that radiation emitted at $Q$ will reach $P$ at time $t$. Thus, the potential at $P$ will depend only on the position described by the vector $a=\overrightarrow{QP}$ of the charge at the proper time $\tilde{\tau}$. See Figure 1. \begin{figure}[h!] \centering \scalebox{0.4}{\includegraphics{pointChargePoten.pdf}} \caption{The four-vectors associated with an observer and a moving charge.}\label{chargePotent} \end{figure} Let $K$ be an inertial reference frame in space-time with coordinates $(ct,x,y,z)=x^\mu,$ where $c$ denotes the speed of light. For the rest of the paper, we will use units in which $c=1$ and omit $c$ from equations. The inner product of two 4-vectors is defined as \begin{equation}\label{inner prod mink} {a}\cdot {b}=\eta _{\mu\nu}a^\mu b^\nu ,\;\;\eta _{\mu\nu}=\mbox{diag}(1,-1,-1,-1). \end{equation} The space of 4-vectors with this inner product is Minkowski space-time $M$. Let $x^\mu$ denote the coordinates of $P$, and let $\tilde{x}^\mu$ be the coordinates of $Q$, the charge at the retarded time. Introduce a 4-vector $a(x)=\overrightarrow{QP}$. Then \begin{equation}\label{a def} a^\mu(x)=x^\mu-\tilde{x}^\mu \quad \mbox{ and } \quad a^2=a\cdot a=0. \end{equation} The vector $a(x)$ is a null (light-like) vector in space-time. Since $a$ is a null vector, we have \begin{equation}\label{null vec relations} (a^0+a^3)(a^0-a^3)=(a^1+ia^2)(a^1-ia^2). \end{equation} We may therefore define a dimensionless complex constant \begin{equation}\label{z in cart} \zeta=\frac{a^1-ia^2}{a^0+a^3}=\frac{a^0-a^3}{a^1+ia^2}. \end{equation} This constant coincides with the ``single complex parameter" occurring during the stereographic projection of the celestial sphere to the Agrand plane (see \cite{PenroseRindler} v.1 p.15). We want the scalar potential to be a function of the dimensionless scalar $\zeta$. To identify the "right" function, note that the electric force depends on the distance from the charge as $\frac{1}{r^2}$ and, as explained in the Introduction, the force is a second derivative of the potential. Hence, the natural candidate for the scalar potential is a multiple of the logarithm function. \vspace{.2in} \noindent\textbf{Definition}\quad We define the \textit{complex potential} or \textit{pre-potential} $S (x)$ at the observer point$x$ of a moving charge $q$ by \begin{equation}\label{WitPoint_1} S(x)=q\ln \zeta=q\ln\frac{a^1(x)-ia^2(x)}{a^0(x)+a^3(x)}, \end{equation} where $a(x)$ is defined by (\ref{a def}). \section{The 4-potential and the Faraday vector of the electromagnetic field} An electromagnetic field can be defined by an electric field intensity $\mathbf{E}(\mathbf{r},t)$ and a magnetic field intensity $\mathbf{B}(\mathbf{r},t)$. Equivalently, one can define a complex 3D-vector $\mathbf{F}$, called the Faraday vector, by \begin{equation}\label{FaradayVector} \mathbf{F}=\mathbf{E}+i\mathbf{B} \end{equation} in order to represent the electromagnetic field. Since $\mathbf{E}$ and $\mathbf{B}$ may be expressed as certain derivatives of the the 4-potential $A=A_\mu$, the Faraday vector $\mathbf{F}$ may also be derived from the 4-potential: \begin{equation}\label{Fj fromA} F_j=2{\partial}^\nu(\rho^ j)^\mu_\nu A_\mu. \end{equation} Here the differential operators are $\partial_\mu=\frac{\partial}{\partial x^\mu}$ and $\partial^\mu=\eta^{\mu\nu}{\partial}_\nu$. The matrices $(\rho^ j)^\mu_\nu$ are the Majorana-Oppenheimer matrices (see \cite{Dvoeglasov}) \[ (\rho^1)_\nu^\mu=\frac{1}{2}\left( \begin{array}{cccc} 0 & 1 & 0 & 0 \\ 1 & 0 & 0 & 0 \\ 0 & 0 & 0 & -i \\ 0 & 0 & i & 0 \\ \end{array}\right),\; (\rho^2)_\nu^\mu=\frac{1}{2}\left( \begin{array}{cccc} 0 & 0 & 1 & 0 \\ 0 & 0 & 0 & i \\ 1 & 0 & 0 & 0 \\ 0 & -i & 0 & 0 \\ \end{array}\right),\] \begin{equation}\label{rho matrices} (\rho^3)_\nu^\mu=\frac{1}{2}\left( \begin{array}{cccc} 0 & 0 & 0 & 1 \\ 0 & 0 & -i & 0 \\ 0 & i & 0 & 0 \\ 1 & 0 & 0 & 0 \\ \end{array}\right)\,. \end{equation} which were used and studied also in \cite{FD}. Direct computation shows that the matrices $\sigma^j=i\rho^j$ obey the commutator relations of the rotation group $SO(3)$, while the $\rho^j$ matrices obey the commutator relations of boosts in the Lorentz group, i.e. \begin{equation}\label{commutationof Kj} [\sigma^j,\sigma^k]=-\epsilon^{jk}_l\sigma^l,\;\;[\rho^j,\rho^k]=\epsilon^{jk}_l\sigma^l\;\; [\sigma^j,\rho^k]=\epsilon^{jk}_l\rho^l\,. \end{equation} As a result, we can use the six matrices $\rho^j,\sigma^j$ as generators of the Lie algebra of the Lorentz group. In addition, the complex conjugates $\bar{\rho}^j$ of these matrices satisfy the same commutation relations. Thus $\bar{\rho}^j$ and $\bar{\sigma}^j=-i\bar{\rho}^j$ also generate the Lie algebra of the Lorentz group. Moreover, \begin{equation}\label{commutation of k kbar} [\bar{\rho}^j,\rho^l]=0\,. \end{equation} In addition to the above commutation relations, these matrices also satisfy the following anti-commutation relations: \begin{equation}\label{CAR} \{\rho^j,\rho^l\}=\rho^j\rho^l+\rho^l\rho^j=\frac{\delta^{jl}}{2}I,\;\; \{\bar{\rho}^j,\bar{\rho}^l\}=\frac{\delta^{jl}}{2}I\,. \end{equation} \section{Lorentz group representations in $\mathcal{M}^4$} Denote by $\mathcal{M}^4$ the complex space $\mathbb{C}^4$ endowed with the bilinear $\mathbb{C}$-valued form $x\cdot y=\eta_{\mu\nu} x^\mu y^\nu$. One possible interpretation of $\mathcal{M}^4$ is the following. Let $\psi:M \rightarrow \mathbb{C}$ be the wave function of a zero spin particle. The gradient operator, describing a generalized momentum, maps space-time into $\mathcal{M}^4$, since $\nabla\psi\in\mathbb{C}^4$. The bilinear form on $\mathcal{M}^4$ is an extension of the inner product on the Minkowski space-time. Our complex scalar potential $S(x)$ is a function $M \rightarrow \mathbb{C}$ as well. Denote by $\pi$ the lift to $\mathcal{M}^4$ of the fundamental representation of the Lorentz group $L$ on $M$, see Figure~\ref{pirep}. \begin{figure}[h!] \scalebox{0.9}{\begin{picture}(500,120)(-125,-20) \put(15,70){\makebox(0,0){$\mathcal{M}^4$}} \put(40,70){\vector(1,0){120}} \put(100,80){\makebox(0,0){$\pi$}} \put(182,70){\makebox(0,0){$\mathcal{M}^4$}} \put(17,15){\vector(0,1){40}} \put(10,35){\makebox(0,0){$\nabla$}} \put(180,15){\vector(0,1){40}} \put(173,35){\makebox(0,0){$\nabla$}} \put(182,1){\makebox(0,0){$M$}} \put(40,1){\vector(1,0){120}} \put(100,12){\makebox(0,0){$\Lambda$}} \put(17,1){\makebox(0,0){$M$}} \end{picture}} \caption{The representation $\pi$}\label{pirep} \end{figure} We denote by $\tilde{\pi}$ the representation in $\mathcal{M}^4$ generated by matrices $\rho^j$ and $i\rho^j$ (a boost in direction $j$ is given by $\Upsilon ^j=\exp (\rho^j)$), and by $\tilde\pi^*$ the representation in $\mathcal{M}^4$ generated by matrices $\bar{\rho}^j$ and $-i\bar{\rho}^j$. Note that the matrices $\rho^j+\bar{\rho}^j$ are generators of boosts in direction $j$. Thus, using (\ref{commutation of k kbar}), the representation $\pi$ of the Lorentz group $L$ can be decomposed as \begin{equation}\label{decomposition of pi boosts} \Lambda ^j=\exp (\rho^j+\bar{\rho}^j)=\exp (\rho^j)\exp (\bar{\rho}^j)=\Upsilon ^j\bar{\Upsilon} ^j\,, \end{equation} or more generally, for any $g\in L$ \begin{equation}\label{decomposition of pi gen} \pi(g)=\tilde{\pi}(g)\tilde\pi^*(g)\,. \end{equation} \section{Covariance under the representations in $\mathcal{M}^4$} In \cite{FD} a \textit{complex Faraday tensor} is introduced for the description of an electromagnetic field, similar to the one introduced by Silberstein \cite{Silberstein}. This tensor is a complex matrix (mixed tensor) $\mathcal{F}^\beta_\alpha =\sum_{j=1}^3 F_j (\rho^j)^\beta_\alpha$, with $F_j$ defined by (\ref{FaradayVector}). We denote its complex conjugate by $\bar{\mathcal{F}}^\beta_\alpha =\sum_{j=1}^3 \bar{F}_j (\bar{\rho}^j)^\beta_\alpha\,.$ With this notation, the usual electromagnetic tensor $F^\beta_\alpha $ can be decomposed as \begin{equation}\label{decomp F} F^\beta_\alpha ={\mathcal{F}}^\beta_\alpha+\bar{\mathcal{F}}^\beta_\alpha\,. \end{equation} We will now prove two claims. \vskip0.2cm\noindent \textbf{Claim 1} \quad The covariance of the tensor $F^\beta_\alpha$ under the representation $\pi$ is equivalent either to the covariance of ${\mathcal{F}}^\beta_\alpha$ under $\tilde{\pi}$ or, equivalently, to the covariance of $\bar{\mathcal{F}}^\beta_\alpha$ under $\tilde{\pi}^*$. \begin{proof} We check the covariance under the boost $ \Lambda ^j$, defined by (\ref{decomposition of pi boosts}), in the direction $j$. Under this transformation, from (\ref{decomp F}) we have \[ F^{\beta'}_{\alpha'}=(\Lambda ^j)^{-1}F^\beta_\alpha\Lambda ^j=(\Upsilon ^j\bar{\Upsilon} ^j)^{-1}( {\mathcal{F}}+\bar{\mathcal{F}})\Upsilon ^j\bar{\Upsilon} ^j.\] Using (\ref{commutation of k kbar}), we get $[\Upsilon ^j,\bar{\rho}^l]=[\Upsilon ^j,\bar{\Upsilon}^l]= [\Upsilon ^j,\bar{\mathcal{F}}]=[\bar{\Upsilon} ^j,\rho^l]=[\bar{\Upsilon} ^j,\mathcal{F}]=0$. Hence, the above equation can be rewritten as \[(\Lambda ^j)^{-1}F^\beta_\alpha\Lambda ^j=(\Upsilon ^j)^{-1}{\mathcal{F}}\Upsilon ^j+(\bar{\Upsilon} ^j)^{-1} \bar{\mathcal{F}}\bar{\Upsilon} ^j.\] This proves Claim 1 for boosts. Similarly, one can establish covariance under action of an arbitrary element of the group $L$. \end{proof} \vskip0.2cm\noindent\textbf{Claim 2} \quad The dimensionless constant $\zeta$ defined by (\ref{z in cart}) and the complex potential $S(x)$ defied by (\ref{WitPoint_1}) are covariant under the representation $\tilde{\pi}$. \begin{proof} Note that from (\ref{CAR}), it follows that $\Upsilon ^j=\exp(\rho^j\psi)= \cosh(\psi/2)I+\sinh(\psi/2)2\rho^j.$ Thus, if we apply, for example, $\Upsilon ^1$ on the vector $a$, we get \[\Upsilon ^1 a=\cosh(\psi/2)(a^0,a^1,a^2,a^3)+\sinh(\psi/2)(a^1,a^0,-ia^3,ia^2).\] So, applying this transformation to $\zeta$ and using the identity in (\ref{z in cart}), we get \[\Upsilon ^1 (\zeta)=\frac{\cosh(\psi/2)(a^1-ia^2)+\sinh(\psi/2)(a^0-a^3)} {\cosh(\psi/2)(a^0+a^3)+\sinh(\psi/2)(a^1+ia^2)}=\] \[\frac{a^1-ia^2}{a^0+a^3}\cdot \frac{1+\tanh(\psi/2)(a^0-a^3)/(a^1-ia^2)} {1+\tanh(\psi/2)(a^1+ia^2)/(a^0+a^3)}=\zeta.\] This proves Claim 2. \end{proof} \section{The Faraday vector and the complex potential of a uniformly moving charge} To define the 4-potential $A$ and consequently the field strength $\mathbf{F}$, we need a new operation on $\mathcal{M}^4$. This operation acts by multiplication with the matrix $C=2\bar\rho ^3$, namely, $a^\mu\mapsto C^\mu_\nu a^\nu$. Since the square of this operation is the identity, we call it the \textit{conjugation}. From (\ref{commutation of k kbar}) it follows that this conjugation is covariant under the representation $\tilde{\pi}$. Define the complex 4-potential $A$ as the conjugate of the gradient of $S$, i.e. \[A_\mu=C_\mu^\lambda{\partial}_\lambda S.\] The Faraday vector $\mathbf{F}$ can be derived from the complex potential by use of (\ref{Fj fromA}): \begin{equation}\label{F from S} F_j={\partial}^\nu (\rho^j)^{\mu}_\nu C_\mu^\lambda{\partial}_\lambda S. \end{equation} This gives explicit formulas for each component of the Faraday vector $F$: \begin{equation}\label{F1 from S} F_1=S_{,13}+iS_{,02},\;\; F_2=S_{,23}-iS_{,01} \end{equation} and \begin{equation}\label{F3 from S} F_3=\frac{1}{2} (S_{,00}-S_{,11}-S_{,22}+S_{,33}) \,. \end{equation} By the above two claims, equation (\ref{F from S}) is covariant under the representation $\tilde{\pi}$. Hence, we will compare this formula with known results only in the case of a rest charge at the origin. Consider a rest charge at the origin. In this case, the world-line of the charge is $L=(t,0,0,0)$. From definition (\ref{a def}), we get $a=(|x|,x^1,x^2,x^3)$, where $|x|=\sqrt{(x^1)^2+(x^2)^2+(x^3)^2}$. Thus, \[S(x)=q\ln\frac{x^1-ix^2}{|x|+x^3}\,.\] Let $\varrho =(x^1)^2+(x^2)^2$. Then, since $\frac{\partial}{\partial x^j}|x|=\frac{x^j}{|x|}$, we obtain $S_{,0}=0$, \[S_{,1}=q\left(\frac{1}{x^1-ix^2}-\frac{x^1/|x|}{|x|+x^3} \right)=\frac{q}{\varrho}\left(x^1+ix^2-\frac{(|x|-x^3)x^1}{|x|}\right)= \frac{q}{\varrho}\left(\frac{x^1x^3}{|x|}+ix^2\right),\] \[S_{,2}=\frac{q}{\varrho}\left(\frac{x^2x^3}{|x|}-ix^1\right),\;\; S_{,3}=-\frac{q}{|x|}\,.\] Then, from (\ref{F1 from S}), we obtain \[ F_1=S_{,13}+iS_{,02}=\frac{\partial}{\partial x^1}S_{,3}=-\frac{\partial}{\partial x^1} \frac{q}{|x|}=\frac{q x^1}{|x|^3}\] and \[ F_2=S_{,23}-iS_{,01}=\frac{\partial}{\partial x^2}S_{,3}=-\frac{\partial}{\partial x^2} \frac{q}{|x|}=\frac{q x^2}{|x|^3}\,.\] To calculate $F_3$ using (\ref{F3 from S}), we first calculate \[S_{,11}=\frac{\partial}{\partial x^1}\frac{q}{\varrho} \left(\frac{x^1x^3}{|x|}+ix^2\right)=-\frac{2qx^1}{\varrho}\left(\frac{x^1x^3}{|x|}+ix^2\right)+ \frac{q}{\varrho^2}\frac{x^3|x|-x^1x^3\frac{x^1}{|x|}}{|x|^2},\] \[S_{,22}=\frac{\partial}{\partial x^2}\frac{q}{\varrho} \left(\frac{x^2x^3}{|x|}-ix^1\right)=-\frac{2qx^2}{\varrho}\left(\frac{x^2x^3}{|x|}-ix^1\right)+ \frac{q}{\varrho^2}\frac{x^3|x|-x^2x^3\frac{x^2}{|x|}}{|x|^2},\] implying that \[S_{,11}+S_{,22}=-\frac{q x^3}{|x|^3}\,.\] Since $S_{,00}=0$ and $S_{,33}=\frac{q x^3}{|x|^3}$, equation (\ref{F3 from S}) yields \[F_3=\frac{q x^3}{|x|^3}.\] This coincides with the usual formula for the electric force of a rest charge. Note that in this case, our complex potential satisfies the wave equation \begin{equation}\label{wave} \square S =S_{,00}-S_{,11}-S_{,22}-S_{,33}=S_{,00}-\nabla ^2S=0\,. \end{equation} Since the d'Alembertian is covariant, the wave equation holds for a field generated by any uniformly moving charge and more generally for any EM field. For a charge $q$ moving uniformly with 4-velocity $u$, the Faraday vector $\mathbf{F}$ can be calculated by \begin{equation}\label{Fjfinal} F_j=q\frac{a_\mu ({\rho}^j)_\nu^\mu u^\nu}{(a \cdot u)^3}\,, \end{equation} where $({\rho}^j)_\nu^\mu$ are defined by (\ref{rho matrices}). From the above calculations for a rest charge at the origin the equation (\ref{Fjfinal}) holds (in this case $u=(1,0,0,0)$). Since this formula is covariant, it also holds in the case of a uniformly moving charge. Equation (\ref{Fjfinal}) coincides with the usual formula for the field of a moving charge (see, for example, \cite{Jackson} p. 573). \section{The scalar potential for an electromagnetic field} Any electromagnetic field is generated by a collection of moving charges. We may assume that charges close to each other move with velocities that do not vary significantly. The sources of the electromagnetic field may be represented by the charge densities $\sigma (y)$ on the space-time 4-vector $y$. We assume that the potential depends additively on the charges generating the field. Thus, the scalar complex potential of the electromagnetic field is given by \begin{equation}\label{scalPotGenera} S(x) =\int\limits_{K^-(x)}\ln\left(\frac{a^1-ia^2}{a^0+a^3}\right)\sigma(x+a)da, \end{equation} where $K^-(x)$ denotes the backward light-cone at $x$. The operators $\alpha_j:=\rho^j C$ occurring in (\ref{F from S}) satisfy the canonical anti-commutation relations similar to (\ref{CAR}) of Dirac's $\alpha$-matrices. Therefore, for any complex potential $S$, equation (\ref{F from S}) can be rewritten as \begin{equation}\label{F from S alfa} F_j={\partial}^\nu (\alpha_j)_\nu ^\lambda{\partial}_\lambda S\,. \end{equation} In the Newman-Penrose basis of $\mathcal{M}^4$ (also known as Bondi tetrad, see \cite{FGW}), the matrices $\alpha_j$ take the usual form $\left( \begin{array}{cc} \sigma _j & 0 \\ 0 & -\sigma _j \\ \end{array} \right)$ where $\sigma _j$ are the Pauli matrices. Note that the matrices $\rho^j$, which define the representation $\tilde{\pi}$, also satisfy the canonical anti-commutation relations (\ref{CAR}). However, they cannot be completed by a $\beta$ matrix, needed for the Dirac equation. The representation $\tilde{\pi}$ is a representation of pairs of spinors, while the representation $\tilde{\pi}^*$ is a representation of pairs of dotted spinors. If the electromagnetic field sources are $J_\mu=(\rho,-j^1,-j^2,-j^3)$, it can be shown \cite{FMaxwell} that the Maxwell equations become \[ \partial_\alpha (\nabla ^2 S)=C_\alpha^\beta J_\beta,\] for $\alpha=0,1,2,3$, added with the wave equation \[ \square S=0\,.\] \section{Discussion} We introduced a new description of an electromagnetic field by a complex-valued function $S(x)$ (pre-potential) on Minkowski space-time. The advantages of this approach are as follows: \begin{itemize} \item The multiple-valued nature of the pre-potential is a natural expression of the Aharonov-Bohm effect. \item Our approach reduces the degrees of freedom from 4 to 2 in the description of an electromagnetic field. \item It reveals a new connection between the Dirac equation and classical electrodynamics. \item We obtain a new complex Lorentz invariant $\zeta$ associated with any null-vector. \item Our approach reveals a new connection between the fundamental and spinor representations of the Lorentz group. \end{itemize} Our future steps are:\begin{itemize} \item Incorporate the pre-potential into the Dirac and Schr\"odinger equations. \item Understand the effect of the electromagnetic field on the solutions of these equations through the pre-potential. \item Derive the formulae for the pre-potential for standard sources of an electromagnetic field. \end{itemize} \ack We would like to thank T. Scarr for editorial comments. \section*{References}
\section{The model} In the present communication we introduce a new kernel which somehow interpolates the Dirac delta and the exponential kernel having the main properties of both. The $ Q(s) = 1/s^l$ is such a function: it is singular at the origin and has a short range of decay for $ l>1$. Let's consider the following relaxation kernel \begin{equation}\label{6} Q(t-t') = \frac{k\tau^l}{(t-t'+\omega)^l} \end{equation} where $k$ is the effective thermal conductivity, $\tau$ is a relaxation time and $l>1$ is a parameter, $-t' + \omega$ is just a time shift which is needed to regularize the expression. Using the general form of heat flux (\ref{1}) we get \begin{equation}\label{7} q = -\int_{-\infty}^{t} \frac{k\tau^l}{(t-t'+\omega)^l} \frac{\partial T(x,t)}{ \partial x}dt'. \end{equation} One has \begin{eqnarray}\label{8} \frac{ \partial q}{\partial t} = -k\left(\frac{\tau}{\omega}\right)^l \frac{\partial T(x,t)}{\partial x} + \nonumber \\ l\int_{-\infty}^{t} \frac{1}{t-t'+\omega} \frac{k\tau^l}{(t-t'+\omega)^l} \frac{ \partial T(x,t')}{\partial x} dt'. \end{eqnarray} A {\it formal} application of integral mean theorem to the second term on the right hand side and the definition of $q$ leads to a new phenomenological law \begin{equation}\label{9} \frac{ \partial q}{\partial t} = -k\left(\frac{\tau}{\omega}\right)^l \frac{ \partial T(x,t)}{\partial x} - \frac{l}{t-t''+\omega} q. \end{equation} The additional energy conservation law is still (\ref{4}) and from this and the last equation we obtain \begin{eqnarray}\label{10} \frac{\gamma}{k} \left( \frac{\omega}{\tau} \right)^l \frac{\partial^2 T(x,t)}{\partial t^2} + \nonumber \\ \frac{\gamma}{k} \left( \frac{\omega}{\tau} \right)^l \frac{l} {t-t''+\omega} \frac{\partial T(x,t)}{\partial t} = \frac{\partial^2 T(x,t)}{\partial x^2}. \end{eqnarray} For a better transparency let's call $\epsilon = \frac{\gamma}{k} \left(\frac{\omega}{\tau}\right)^l $ and $a = \frac{\gamma}{k} \left(\frac{\omega}{\tau}\right)^l \cdot l$ The physical meaning of $\epsilon$ is still the thermal diffusivity multiplied by a scaling constant which is the renormalized relaxation time (the ratio of an ordinary time shift $\omega$ and a well defined relaxation time $\tau$). The exponential $l$ is a real number which describes the non-locality in time which we may call memory effects of the heat conduction phenomena. Larger $l$ means shorter memory. The physical meaning of $a$ is approximately the thermal diffusivity multiplied by another time-scaling factor. In the following we will see that the role of $a, \epsilon$ or $l$ will be crucial in the structure of the solutions. At last we introduce a new time variable $ t = t-t''+ \omega.$ Now our telegraph-type equation reads \begin{equation}\label{11} \epsilon\frac{\partial^2 T(x,t)}{\partial t^2} + \frac{a}{t} \frac{\partial T(x,t)}{\partial t} = \frac{\partial^2 T(x,t)}{\partial x^2}. \end{equation} Note, that the $a/t$ factor appearing in front of the first time derivate makes the equation time-reversible, which cannot be true for diffusion or heat propagation processes, at the same time the $a/t$ factor makes the equation irregular at the origin. To avoid these problems, we may shift the pole to a negative time value - in practical calculations using the $a/(t+\tau)$ where $\tau$ still can be any kind of relaxation time with well-founded physical interpretation. Physically it is clear, if a process has a well-defined time-scale than the reverse process cannot run back in time more than the physically relevant time. Now, in this sense, for positive time $(t>0)$ we may use the equation to describe diffusion-like processes. Our deeper investigation clearly shows, that no other exponent of $t$ as one can have self-similar solution. If we consider (\ref{11}) as a non-linear wave equation we may investigate wave properties like dispersion phenomena. Inserting the standard plain wave approximation $T(x,t) = e^{i(\tilde{k}x + \tilde{\omega}t)}$ into (\ref{11}) the dispersion relation and the attenuation distance can be obtained. These are the followings: \begin{equation} v_p = \frac{\tilde{\omega}}{Re(\tilde{k})} = \sqrt{\frac{2}{\epsilon}} {\tilde{\omega}} \left( 1+ \sqrt{1 + \left( \frac{l}{t} \right)^2 } \right )^{-1/2} \hspace{0.3cm} \tilde{\alpha} = \frac{1}{Im(\tilde{k})} = \frac{2t}{\epsilon l}\frac{1}{v_p} . \end{equation} Our telegraph-type equation is time dependent, hence both the dispersion relation and the attenuation distance have time dependence. Note, that $v_p$ has a very weak time-dependence, basically only till $t \le l$. As an other interesting point is that the phase velocity does not depend on the angular velocity which is the same as for the ideal wave equation, so our equation has no dispersion. So in this sense our equation is very similar to the wave-equation, which is hyperbolic. The properties of the attenuation distance is even more interesting, it is divergent in time and has a $1/\tilde{\omega}$ angular frequency. However if we let the angular frequency and the time to go infinite with the same speed than the attenuation distance has a strong decay. Which is like the skin-effect when high frequency electrons can only propagate on the surface of a metal. We are looking for solution of (\ref{11}) of the form \begin{equation} T(x,t)=t^{-\alpha}f\left(\frac{x}{t^\beta}\right):=t^{-\alpha}f(\eta). \end{equation} The similarity exponents $\alpha$ and $\beta$ are of primary physical importance since $\alpha$ represents the rate of decay of the magnitude $T(x,t)$, while $\beta$ is the rate of spread (or contraction if $\beta<0$ ) of the space distribution as time goes on. Substituting this into (\ref{11}) we have \begin{eqnarray}\label{12} f''(\eta) t^{-\alpha -2} [\epsilon\beta^2 \eta^2] + \nonumber \\ f'(\eta) \eta t^{-\alpha -2} [\epsilon\alpha\beta - \epsilon\beta(-\alpha-\beta-1) - \beta a] + \nonumber \\ f(\eta) t^{-\alpha -2} [-\epsilon\alpha(-\alpha-1) - a\alpha] = f''(\eta) t^{-\alpha -2\beta} \label{3}, \end{eqnarray} where prime denotes differentiation with respect to $\eta.$ One can see that this is an ordinary differential equation(ODE) if and only if $\alpha + 2 = \alpha + 2\beta$ ({\it {the universality relation}}). So it has to be \begin{equation} \beta =1 \end{equation} while $\alpha$ can be any number. The corresponding ODE we shall deal with is \begin{equation}\label{13} f''(\eta) [\epsilon\eta^2-1 ] + f'(\eta) \eta(2\epsilon\alpha + 2\epsilon -a) + f(\eta) \alpha (\epsilon \alpha+\epsilon- a) = 0. \end{equation} In pure heat conduction-diffusion processes--no sources or sinks--the heat mass is conserved: the integral of $T(x,t)$ with respect to $x$ does not depend on time $t$. For $T(x,t)$ this means \begin{equation} \int T(x,t) dx=t^{-\alpha} \int f(\frac{x}{t})dx=t^{-\alpha +1}\int f(\eta)d\eta=const \end{equation} if and only if $\alpha=1.$ We are going to investigate this case only. Plainly (\ref{14}) can be written as \begin{equation} (\varepsilon f\eta^2-f)''=a(\eta f)' \end{equation} which after integration and supposing $f(\eta_0)=0$ for some $\eta_0$ gives \begin{equation}\label{14} \frac{df}{f}=\frac{a\eta d\eta}{\varepsilon \eta^2-1}. \end{equation} From this equation we can obtain two qualitatively different solutions. The one which is globally bounded and positive in the domain $\{(x,t): 1-\varepsilon \eta^2 >0 \}$ and has the form \begin{equation} f=(1-\varepsilon \eta^2)^{\frac{a}{2\varepsilon}-1}_+ \end{equation} where $(f)_+=\max (f,0).$ See Fig. (\ref{elso}). The corresponding self-similar solution is \begin{equation}\label{15} T(x,t)=\frac{1}{t}\left(1-\varepsilon \frac{x^2}{t^2}\right)^{\frac{a}{2\varepsilon}-1}_+ \end{equation} This solution is positive in the cone $t^2> \varepsilon x^2$ and is zero outside of it, see Fig. (\ref{kettes}). Note, that only the $x>0$ and $t>0$ quarter of the plane is presented, because of it has physical relevance. \begin{figure}* \scalebox{0.65}{ \rotatebox{0}{\includegraphics{elso.eps}}} \vspace*{0.4cm} \caption{Eq. (21) thick solid line is for $l = 6.2$ and the thin dashed line is for $l = 4.1$. } \label{elso} \end{figure} On the $(x,t)$ plane there are two fronts $x(t)=\pm \frac{t}{\sqrt{\varepsilon}}$ separating these domains. Because the function $T(x,t)$ not always has continuous derivatives entering to (\ref{11}) we have to make clear what we mean under "solution". Having in mind the physical background, we ask the continuity of $T_t, T_x, q_t$ and $q_x$ so that in (2) and (3) all functions were continuous. \begin{figure}* \scalebox{0.35}{ \rotatebox{-90}{\includegraphics{Fig2_f.eps}}} \caption{The solution (\ref{15}) for the parameter $l=6.2$. } \label{kettes} \end{figure} In our case this means that \begin{equation}\label{16} \frac{a}{2\varepsilon}-1=\frac{l-2}{2} \end{equation} has to be greater than $1$, i.e. $a/\varepsilon=l>4,$ which we shall suppose further on. If the second derivatives are not continuous ($4<l\leq 6$) we understand the solution in the sense of distributions. If $l>6$ the solution is classical. On Fig. 1. we compare the solutions with $l=4.1$ and $l=6.2$. the thick solid line represents the solution for $ l= 6.2$ and the thin dashed line however shows the solution for $l=4.1$. {\bf Remark 1.} The solution (\ref{15}) is of {\it source-type} i.e. $\lim_{t\rightarrow 0}T(x,t)= K\delta(x),$ where $\delta$ is the Dirac measure, $K>0$. One can calculate the second initial condition $\lim_{t\rightarrow 0}T_t(x,t)$ too. {\bf Remark 2.} One can write (\ref{15}) in the form of {\it product} of two traveling waves propagating in opposite direction (divided by a time-factor): \begin{equation} T(x,t)=\frac{1}{t^{l-1}}(t-\sqrt{\varepsilon}x)_+^{\frac{l}{2}-1}(t+\sqrt{\varepsilon}x)_+^{\frac{l}{2}-1} \end{equation} which is a new-type of purely hyperbolic wave; the typical solution of the wave equation is the {\it sum} of two such waves: $g(x-ct)+g(x+ct)$. It is known that an another possible answer to contradiction connected with the infinite speed of propagation is the nonlinear Fourier law ($\tau=0, k=k_0 T^{m-1}$ in (2)) which leads to a nonlinear heat equation \begin{equation}\label{17} T_t=(T^m)_{xx}, \quad m>1. \end{equation} In \cite{zk} Zeldovich and Kompaneets have found the fundamental solution $T_1$ of this equation which we write in the following form: \begin{eqnarray}\label{18} T_1^{m-1}=t^{-\alpha (m-1)}\left(A^2-B^2x^2t^{-2\beta}\right)_+= \nonumber \\ \frac{1}{t}(At^\beta-Bx)_+(At^\beta+Bx)_+, \end{eqnarray} where $A$ is constant and \begin{equation} \alpha=\beta=\frac{1}{m+1},\quad B^2=\frac{m-1}{2m(m+1)}. \end{equation} One can see that this solution has bounded support in $x$ for any $t>0$ which is a hyperbolic property. Using comparison principle for such equations one can show this finite speed property for any initial condition having compact support. However, the fronts are not straight lines: $x(t)=\pm \frac{A}{B}t^\beta$, $\beta <1$ so the speed of propagation $\dot{x}(t)$ goes to zero if $t$ goes to infinity. One can also see that $T_1$ is of source-type: $T_1(x,0)=K_1\delta(x).$ The most intrinsic property of $T_1$ is that it plays the role of {\it intermediate asymptotic}: any solution of (\ref{17}) corresponding to the initial datum $t(x,0)$ with $\int t(x,0)dx=K_1$ converges to $T_1$ as $t\rightarrow \infty.$ This was conjectured earlier but was shown only in 1973 by Sh. Kamin, see \cite{ka}. It would be important and interesting to understand whether or not our special solution $T(x,t)$ had this attractor property. If "yes", in what sense: we recall that there is a second initial condition too. \\ {\it{In summary. -}} We introduced a new phenomenological law for heat flux which in some sense "interpolates" between Fourier and Cattaneo laws. The consequence of it is a non-autonomous model, a telegraph-type partial differential equation. It already has, unlike the classical telegraph equation, self-similar solutions, the presence of which is desirable in the theory of heat propagation free from sources and absorbers. One of us (R.K.) would like to thank Prof. P. Rosenau for his stimulating discussion.
\section{The conjecture of Birch and Swinnerton-Dyer and the parity conjecture} Let $K$ be a number field and $E$ an elliptic curve defined over $K$. Denote by $K_{v}$ the completion of $K$ at a place $v$. We recall a few definitions: \begin{definition} (Tate Module):\newline The $l$-adic Tate module of $E$ is the inverse limit of the system of multiplication by $l$ maps $E[l^{n+1}]\longrightarrow E[l^{n}]$, where $E[m]$ denotes the kernel of multiplication by $m$ on $E.$\newline Set $T_{l}(E)=\lim\limits_{\longleftarrow }E[l^{n}]$, $V_{l}(E)=\mathbb{Q}% _{l}\otimes _{\mathbb{Z}_{l}}T_{l}(E)$ and:% \begin{equation*} \sigma _{E/K_{v},l}^{\prime }:Gal(\overline{K}_{v}/K_{v})\longrightarrow GL(V_{l}(E)^{\ast }). \end{equation*} \end{definition} \begin{definition} Fix an embedding, $\iota :\mathbb{Q}_{l}\hookrightarrow \mathbb{C}$; we can then associate to $\sigma _{E/K_{v},l}^{\prime }$ a complex representation $% \sigma _{E/K_{v},l,\iota }^{\prime }$ of the Weil-Deligne group (see \cite% {Roh1} \S 13).\newline One can show that the isomorphism class of $\sigma _{E/K_{v}}^{\prime }:=\sigma _{E/K_{v},l,\iota }^{\prime }$ is independent of the choice of $l$ and $\iota $ (see \cite{Roh1} \S 13, \S 14, \S 15). \end{definition} Denote by $L(E/K,s)$ the global $L$-function, product of local $L$-functions:% \begin{equation*} L(E/K,s)=\tprod\limits_{v\text{ finite}}L(E/K_{v},s)\left( =\tprod\limits_{v% \text{ finite}}L(\sigma _{E/K_{v}}^{\prime },s)\right) , \end{equation*}% defined for $\func{Re}(s)>\dfrac{3}{2}$ (see \cite{Roh1} \S 17 for the correspondence between the classical definition of $L(E/K_{v},s)$ and the one using $\sigma _{E/K_{v}}^{\prime }$) and by \begin{equation*} \Lambda (E/K,s)=A(E/K)^{s/2}L(E/K,s)(2(2\pi )^{-s}\Gamma (s))^{[K:\mathbb{Q}% ]}\text{,} \end{equation*}% the "complete" $L$-function. Recall the following classical conjectures: \begin{conjecture} (Birch and Swinnerton-Dyer: BSD):\newline $ord_{s=1}\Lambda (E/K,s)=rk(E/K).$ \end{conjecture} \begin{conjecture} \label{conj eq funct}(Functional equation of $\Lambda :$ FE):\newline $L(E/K,s)$ has a holomorphic continuation to $\mathbb{C}$ and there is a number\newline $W(E/K)=\tprod\limits_{v}W(E/K_{v})\in \{\pm 1\}$ such that:% \begin{equation*} \Lambda (E/K,s)=W(E/K)\Lambda (E/K,2-s) \end{equation*}% (see \cite{Roh1} \S 12 and \S 19 for the definition of $W(E/K_{v}):=W(\sigma _{E/K_{v}}^{\prime })$ and \cite{Roh1} \S 21 p.157 for the functional equation of $\Lambda $). \end{conjecture} This conjecture is known in a few cases: \begin{itemize} \item For elliptic curves over $\mathbb{Q}$ thanks to modularity results on elliptic curves due to Wiles, Taylor, Breuil, Diamond and Conrad \item For elliptic curves over a totally real field $K$, we only know a meromorphic continuation and the functional equation of $\Lambda $ thanks to a potential modularity result of Wintenberger (see \cite{Wint}) together with an argument of Taylor. \end{itemize} In general, conjecture \ref{conj eq funct} is not known. The conjecture of Birch and Swinnerton-Dyer implies the following weaker conjecture: \begin{conjecture} (BSD $\left( \func{mod}2\right) $)\newline $\func{rk}(E/K)\equiv ord_{s=1}\Lambda (E/K,s)$ $\left( \func{mod}2\right) .$ \end{conjecture} Combining it with the conjectural functional equation we get: \begin{conjecture} \label{conj parity}(Parity conjecture)\newline $(-1)^{\func{rk}(E/K)}=W(E/K).$ \end{conjecture} Tim and Vladimir Dokchitser showed that this conjecture is true assuming that the $6^{\infty }$-part of the Tate-Shafarevich group of $E$ over $% K(E[2])$ is finite (see \cite{Dok3} Th 7.1 p.20). \begin{definition} Selmer group:\newline Let $X_{p}(E/K):=Hom_{\mathbb{Z}_{p}}(S(E/K,p^{\infty }),\mathbb{Q}_{p}/% \mathbb{Z}_{p})\otimes _{\mathbb{Z}_{p}}\mathbb{Q}_{p},$\newline where $S(E/K,p^{\infty }):=\lim\limits_{\underset{n}{\longrightarrow }% }S(E/K,p^{n})$ is the $p^{\infty }$-Selmer group, sitting in an exact sequence:% \begin{equation*} 0\longrightarrow E(K)\otimes \mathbb{Q}_{p}/\mathbb{Z}_{p}\longrightarrow S(E/K,p^{\infty })\longrightarrow \Sha_{E/K}[p^{\infty }]\longrightarrow 0 \end{equation*} \end{definition} If we let $\func{rk}_{p}(E/K):=\dim _{_{\mathbb{Q}_{p}}}X_{p}(E/K)=\func{rk}% (E/K)+\func{cork}_{\mathbb{Z}_{p}}\Sha_{E/K}[p^{\infty }]$, a more accessible form of the conjecture \ref{conj parity} is the following: \begin{conjecture} \label{conj p-parity}($p$-parity conjecture)\newline $(-1)^{\func{rk}_{p}(E/K)}=W(E/K)$. \end{conjecture} If $L/K$ is a finite Galois extension and $\tau $ is a self-dual $\overline{% \mathbb{Q}}_{p}$-representation of $Gal(L/K)$ then there is an equivariant form of the conjecture \ref{conj p-parity}: \begin{conjecture} ($p$-parity conjecture for (self-dual) twists)\newline $(-1)^{\left\langle \tau ,X_{p}(E/K)\right\rangle }=W(E/K,\tau )$, where $% W(E/K,\tau )=\tprod\limits_{v}W(\sigma _{E/K_{v}}^{\prime }\otimes \limfunc{% Res}\nolimits_{D_{v}}\tau )$\newline (where $D_{v}\subset Gal(L/K)$ is the decomposition group at $v$). \end{conjecture} It is this last conjecture in a particular setting that will interest us for the rest of the paper. \section{Invariance of the parity conjecture in a $D_{2p^{n}}$-extension \label{section}} \subsection{Statement of the main theorem and applications to the $p$-parity conjecture} Let $K$ be a number field, $E/K$ an elliptic curve and $L/K$ a finite Galois extension such that $Gal(L/K)\simeq D_{2p^{n}},$ with $p\geq 5$ a prime number. $D_{2p^{n}}$ admits the following irreducible representations over $% \overline{\mathbb{Q}}_{p}$:\newline $\bullet $ $1$ the trivial representation\newline $\bullet $ $\eta $ the quadratic character\newline $\bullet $ $\frac{p^{n}-1}{2}$ irreducible representations of degree 2; they are of the form,% \begin{equation*} I(\chi ):=Ind_{C_{p^{n}}}^{D_{2p^{n}}}(\chi )=I(\chi ^{-1}), \end{equation*}% where $\chi $ is a non-trivial character of $C_{p^{n}}$ ($I(1)=1\oplus \eta $ is reducible)$.$\newline See for example \cite{Ser1} for the description of irreducible representations of $D_{2p^{n}}$. \medskip Let $\tau =I(\chi )$ be such an irreducible representation of degree 2. \begin{theorem} \label{theo2}With the notation above and $p\geq 5$, we have the following equality:% \begin{equation*} \frac{W(E/K,\tau )}{W(E/K,1\oplus \eta )}=\dfrac{(-1)^{\left\langle \tau ,X_{p}(E/L)\right\rangle }}{(-1)^{\left\langle 1\oplus \eta ,X_{p}(E/L)\right\rangle }} \end{equation*}% In other words, the $p$-parity conjecture for $E/K$ tensored by $1\oplus \eta \oplus \tau $ holds:% \begin{equation*} W(E/K,1\oplus \eta \oplus \tau )=(-1)^{\left\langle 1\oplus \eta \oplus \tau ,X_{p}(E/L)\right\rangle } \end{equation*} \end{theorem} \begin{remark} \label{cas des freres Dok}The Dokchitser brothers have shown that this equality holds in two different cases:\newline $\bullet $ \textit{In the case when }$p$\textit{\ is any prime number but the elliptic curve }$E/K$\textit{\ has a cyclic decomposition group in all additive places above }$2$\textit{\ and }$3$ (see\textit{\ }\cite{Dok2} Th.4.2 (1) p.43).\textit{\smallskip }\newline $\bullet $ \textit{In the case when }$p\equiv 3\mathit{\ }\left( \func{mod}% 4\right) $\textit{\ (without any additional assumption) using a strong global }$p$-\textit{parity result over totally real fields }(see\textit{\ }% \cite{Dok3} Prop.6.12 p.18).\textit{\smallskip }\newline In particular, the statement of Thm.\ref{theo2} also holds for $p=3.$ Futhermore, this case can be proved without using the "painful calculation" (of \cite{Dok2} p.30) in the case of additive reduction (see the appendix).% \textit{\medskip }\newline \textit{Here we prove the equality for all }$p\geq 5$\textit{\ (without any additional assumption).} \end{remark} \begin{corollary} \begin{equation*} \dfrac{W(E/K,I(\chi ))}{(-1)^{\left\langle I(\chi ),X_{p}(E/L)\right\rangle }% }\text{ does not depend on }\chi :C_{p^{n}}\longrightarrow \mathbb{C}^{\ast }. \end{equation*} \end{corollary} Theorem \ref{theo2} is equivalent to the fact that Hypothesis 4.1 of \cite% {Dok2} holds for any elliptic curve and any $p>3$ (using a result of the Dokchitser brothers it is also true for $p=3$, see Remark \ref{cas des freres Dok} above). Now using the machinery of the Dokchitser brothers (see Th.4.3 and Th.4.5 in \cite{Dok2}) we have the following theorems: \begin{theorem} Let $K$ be a number field, $p\geq 3,$ and $E/K$ an elliptic curve. Suppose $% F $ is a $p$-extension of a Galois extension $M/K,$ Galois over $K.\ $If the $p $-parity conjecture% \begin{equation*} (-1)^{\func{rk}_{p}E/L}=W(E/L) \end{equation*}% holds for all subfields $K\subset L\subset M,$ then it holds for all subfields $K\subset L\subset F.$ \end{theorem} \begin{theorem} Let $K$ be a number field, $p\geq 3,$ $E/K$ an elliptic curve and $F/K$ a Galois extension. Assume that the $p$-Sylow subgroup $P$ of $G=Gal(F/K)$ is normal and $G/P$ is abelian. If the $p$-parity conjecture holds for $E$ over $K$ and its quadratic extensions in $F,$ then it holds for all twists of $E$ by orthogonal representations of $G.$ \end{theorem} \subsection{Reduction to the case of a $D_{2p}$-extension} Here we reduce the demonstration of Theorem \ref{theo2} by an induction argument together with the Galois invariance of root numbers due to Rohrlich (see \cite{Roh3} Theorem 2), to the following statement:\medskip \begin{proposition} It is sufficient to prove Theorem \ref{theo2} in the case when $n=1$ (i.e. $% Gal(L/K)\simeq D_{2p}$). \end{proposition} \begin{proof} Suppose Theorem \ref{theo2} is true for $n=N-1.$ We will show that theorem is true for $n=N$.\newline Consider $L/K$ a finite Galois extension such that $Gal(L/K)\simeq D_{2p^{N}} $ and $\tau $ an irreducible representation of degree $2$ of $% D_{2p^{N}}.$\newline $\bullet $ If $\chi $ is not injective, then the statement is known by the induction hypothesis.\newline $\bullet $ If $\chi $ is injective:\newline Let $\sigma =res(I(\chi )):=res_{D_{2p^{N-1}}}^{D_{2p^{N}}}(I(\chi ))$.% \newline Then $\sigma =I(\chi ^{\prime })$, where $\chi ^{\prime }:=\chi _{\left\vert C_{p^{N-1}}\right. }:C_{p^{N-1}}\rightarrow \overline{\mathbb{Q}}_{p}$ is injective.\newline We have: $Ind_{D_{2p^{N-1}}}^{D_{2p^{N}}}(\sigma )=\tbigoplus\limits_{\chi _{0}}I(\chi _{0})$, where the sum is taken over the $\chi _{0}$ such that $% \chi _{0\left\vert C_{p^{N-1}}\right. }=\chi _{\left\vert C_{p^{N-1}}\right. }.$\newline For each such $\chi _{0}$ there is an element of $Aut(\mathbb{C)}$ sending $% \chi $ into $\chi _{0}$ and $I(\chi )$ into $I(\chi _{0}).$\newline By inductivity of root numbers in Galois extension:% \begin{equation*} W(E/K,\sigma )=W(E/K,Ind_{D_{2p^{N-1}}}^{D_{2p^{N}}}(\sigma )). \end{equation*}% By Galois invariance of root numbers:% \begin{equation*} W(E/K,I(\chi ^{\prime }))=W(E/K,I(\chi _{0}))\text{, }\forall \chi _{0}\text{ such that }\chi _{0\left\vert C_{p^{N-1}}\right. }=\chi _{\left\vert C_{p^{N-1}}\right. }. \end{equation*}% So $W(E/K,\sigma )=W(E/K,Ind_{D_{2p^{N-1}}}^{D_{2p^{N}}}(\sigma ))=W(E/K,\tau )^{p}=W(E/K,\tau ).$\newline On the other hand,\newline $\left\langle \sigma ,X_{p}(E/L)\right\rangle =\left\langle Ind_{D_{2p^{N-1}}}^{D_{2p^{N}}}(\sigma ),X_{p}(E/L)\right\rangle =p.\left\langle \tau ,X_{p}(E/L)\right\rangle $,\newline because $X_{p}(E/L)$ is a $\mathbb{Q}_{p}$-representation$.$\newline So $(-1)^{\left\langle 1\oplus \eta \oplus \sigma ,X_{p}(E/L)\right\rangle }=(-1)^{\left\langle 1\oplus \eta \oplus \tau ,X_{p}(E/L)\right\rangle }.$% \newline By the induction hypothesis, $(-1)^{\left\langle 1\oplus \eta \oplus \sigma ,X_{p}(E/L)\right\rangle }=W(E/K,\sigma )$. As a result,\newline $W(E/K,1\oplus \eta \oplus \tau )=(-1)^{\left\langle 1\oplus \eta \oplus \tau ,X_{p}(E/L)\right\rangle }.$ \end{proof} \subsection{The case of a $D_{2p}$-extension} We first restate Theorem \ref{theo2} in the case of a $D_{2p}$-extension. Let $K$ be a number field, $E/K$ an elliptic curve and $L/K$ a Galois extension such that $Gal(L/K)\simeq D_{2p}\simeq C_{p}\rtimes C_{2},$ with $% p\geq 5$ a prime number. Recall the irreducible representations of $D_{2p}$ over $\overline{\mathbb{Q}% }_{p}$:\newline $\bullet $ $1$ the trivial representation\newline $\bullet $ $\eta $ the quadratic character\newline $\bullet $ $I(\chi )$ irreducible representations of degree 2, where $\chi $ is a non-trivial character of $C_{p}$. \begin{theorem} \label{theo}With the notation above and $p\geq 5$, we have the following equality:% \begin{equation*} \frac{W(E/K,\tau )}{W(E/K,1\oplus \eta )}=\dfrac{(-1)^{\left\langle \tau ,X_{p}(E/L)\right\rangle }}{(-1)^{\left\langle 1\oplus \eta ,X_{p}(E/L)\right\rangle }}. \end{equation*}% In other words, the $p$-parity conjecture for $E/K$ tensored by $1\oplus \eta \oplus \tau $ holds:% \begin{equation*} W(E/K,1\oplus \eta \oplus \tau )=(-1)^{\left\langle 1\oplus \eta \oplus \tau ,X_{p}(E/L)\right\rangle }. \end{equation*} \end{theorem} The proof of Theorem \ref{theo} will occupy the rest of section \ref{section}% . \bigskip We use the following notation:\newline $\bullet $ $v$ a finite place of $K$\newline $\bullet $ $K_{v}$ the completion of $K$ at $v$\newline $\bullet $ $q_{v}=l_{v}^{r}$ the cardinality of the residue field of $K_{v}$% \newline $\bullet $ $z$ $\mid v$ a finite place of $L$\newline $\bullet $ $w\mid v$ a finite place of $L^{H}$ (where $H$ is a subgroup of $% Gal(L/K)=D_{2p}$)\newline $\bullet $ $\delta =ord_{v}\left( \text{the minimal discriminant of }% E/K_{v}\right) $\newline $\bullet $ $\delta _{H}=ord_{w}\left( \text{the minimal discriminant of }% E/\left( L^{H}\right) _{w}\right) $\newline $\bullet $ $e_{H}$ the ramification index of $\left( L^{H}\right) _{w}/K_{v}$% \newline $\bullet $ $f_{H}$ the residue degree of $\left( L^{H}\right) _{w}/K_{v}$% \newline $\bullet $ $\omega _{E/K_{v}}^{0}=$ a minimal invariant differential of $% E/K_{v}$\newline $\bullet $ $C_{w}(E/L^{H})=c_{w}(E/L^{H})\omega (H),$\newline where $% \begin{cases} c_{w}(E/L^{H})=\text{ local Tamagawa factor of }E/L^{H} \\ \omega (H)=\left\vert \frac{\omega _{E/K_{v}}^{0}}{\omega _{E/\left( L^{H}\right) _{w}}^{0}}\right\vert _{\left( L^{H}\right) _{w}}% \end{cases}% $ Furthermore, if $l_{v}>3$ then:\newline \begin{tabular}{ll} $\left\vert \frac{\omega _{E/K_{v}}^{0}}{\omega _{E/\left( L^{H}\right) _{w}}^{0}}\right\vert _{\left( L^{H}\right) _{w}}$ & $=q^{\frac{\delta .e_{H}-\delta _{H}}{12}f_{H}}$ \\ & $=q^{\left\lfloor \frac{\delta .e_{H}}{12}\right\rfloor f_{H}}$ (in the case of potentially good reduction)% \end{tabular} \medskip For $D_{2p}$, there is the following equality: $Ind_{\left\{ 1\right\} }^{D_{2p}}1-2.Ind_{D_{2}}^{D_{2p}}1-Ind_{C_{p}}^{D_{2p}}1+2.1=0$ of virtual representations of $G$, this gives the $G$-relation $\Theta :\left\{ 1\right\} -2D_{2}-C_{p}+2G$ in the sense of \cite{Dok2} (Def 2.1 p.11). \medskip We recall two definitions in our setting (i.e. with $\Theta :\left\{ 1\right\} -2D_{2}-C_{p}+2D_{2p}$), for general definitions see \cite{Dok2}. \begin{definition} (\cite{Dok2}, Def.2.13 p.14): Let $\rho $ be a self-dual $\mathbb{Q}_{p}[G]$% -representation.\newline Pick a $G$-invariant non-degenerate $\mathbb{Q}_{p}$-linear pairing $% \left\langle ,\right\rangle $ on $\rho $ and set\newline $C_{\Theta }(\rho )=\det (\left\langle ,\right\rangle \left\vert \rho ^{\left\{ 1\right\} }\right. )\det (\tfrac{1}{2}\left\langle ,\right\rangle \left\vert \rho ^{D_{2}}\right. )^{-2}\det (\tfrac{1}{p}\left\langle ,\right\rangle \left\vert \rho ^{C_{p}}\right. )^{-1}\det (\tfrac{1}{2p}% \left\langle ,\right\rangle \left\vert \rho ^{D_{2p}}\right. )^{2}.$\newline As an element of $\mathbb{Q}_{p}^{\ast }/\mathbb{Q}_{p}^{\ast 2}$, this does not depend on the choice of the pairing. \end{definition} \begin{definition} (\cite{Dok2}, Def.2.50 p.23) We define:% \begin{equation*} T_{\Theta ,p}=\left\{ \begin{tabular}{l} $\sigma $ a self-dual $\overline{\mathbb{Q}}_{p}[G]\text{-}$ \\ $\text{representation}$% \end{tabular}% \left\vert \begin{tabular}{l} $\left\langle \sigma ,\rho \right\rangle \equiv ord_{p}C_{\Theta }(\rho )$ $% \left( \func{mod}2\right) $ \\ $\forall \rho \text{ a self-dual }\mathbb{Q}_{p}[G]\text{-representation}$% \end{tabular}% \right. \right\} \end{equation*} \end{definition} \medskip Following the approach of the Dokchitser brothers, we have the following theorem \begin{theorem} (Theorem 1.14 of \cite{Dok2}). Let $L/K$ be a Galois extension of number fields with Galois group $G=D_{2p}$, where $p>2$ is a prime number. Let $% \Theta :\left\{ 1\right\} -2D_{2}-C_{p}+2D_{2p}$. For every elliptic curve $% E/K,$ the $\mathbb{Q}_{p}[G]$-represention $X_{p}(E/L)$ is self-dual, and% \begin{equation*} \forall \sigma \in T_{\Theta ,p},\quad(-1)^{\left\langle \sigma ,X_{p}(E/L)\right\rangle }=(-1)^{ord_{p}(C)}, \end{equation*}% where $C=\tprod\limits_{v\nmid \infty }C_{v}$ with $C_{v}=C_{v}(\{1% \})C_{v}(D_{2})^{-2}C_{v}(C_{p})^{-1}C_{v}(G)^{2}$\newline and $C_{v}(H)=\tprod\limits_{\underset{w\text{ places of }L^{H}}{w\mid v}% }C_{w}(E/L^{H}).$ \end{theorem} Now, since $1\oplus \eta \oplus \tau \in T_{\Theta ,p}$ (see \cite{Dok2}, example 2.53 p.24), we only need to prove that :% \begin{equation*} \frac{W(E/K,\tau )}{W(E/K,1\oplus \eta )}=(-1)^{ord_{p}C}. \leqno (1) \end{equation*} Furthermore, since we are only interested in the parity of $ord_{p}(C),$ we do not have to determine $C_{v}(D_{2})$ and $C_{v}(G),$ because these terms only bring an even contribution (since they appear with an even exponent). Both sides of $(1)$ are of local nature. As $W(E/K,\tau )=\tprod\limits_{v}W(E/K_{v},\tau _{v})$, where $\sigma _{v}:=res_{Gal(L_{z}/K_{v})}\sigma ,$ all we need to do is to prove the following local equality:% \begin{equation*} \frac{W(E/K_{v},\tau _{v})}{W(E/K_{v},(1\oplus \eta )_{v})}% =(-1)^{ord_{p}\left( C_{v}\right) },\leqno(2) \end{equation*}% for each finite place $v$ of $K$ ($v\mid \infty $ do not contribute, since $% p\neq 2$). \bigskip Denote by $G_{v}:=Gal(L_{z}/K_{v})$ the decomposition group of $v$. The proof of Theorem \ref{theo} split in several cases:\medskip \begin{tabular}{ll} ${\bullet }$ $G_{v}=\{1\}$ (there are $2p$ places above $v$ in $L$) & see section \ref{cases 1 et C_p} \\ ${\bullet }$ $G_{v}=D_{2}$ (there are $p$ places above $v$ in $L$) & see section \ref{case D_2} \\ ${\bullet }$ $G_{v}=C_{p}$ (there are $2$ places above $v$ in $L$) & see section \ref{cases 1 et C_p} \\ ${\bullet }$ $G_{v}=D_{2p}$ (there is a unique place above $v$ in $L$) & see section \ref{case D_2p}% \end{tabular} \bigskip We first recall a few facts about the local Tamagawa factors of elliptic curves. \subsubsection{Local Tamagawa factors of elliptic curves\label{tamagawa}} The assumptions and notation from above are in force. The local Tamagawa factor at $v,$ $c(E/K_{v})=\#\left( E(K_{v})/E^{0}(K_{v})\right) ,$\newline $\left( \text{where }E^{0}(K_{v})=\left\{ \text{Points of non-singular reduction}\right\} \right) $ is determined by Tate's algorithm (see \cite% {Silv2} IV \S 9):\medskip $c(E/K_{v})=% \begin{cases} 1 & \begin{array}{c} \text{if }E\text{ has good reduction at }v% \end{array} \\ 1,\text{ }2,\text{ }3\text{ or }4 & \begin{array}{c} \text{if }E\text{ has additive reduction at }v% \end{array} \\ n & \begin{array}{l} \text{if }E\text{ has split multiplicative reduction} \\ \text{of type }I_{n}\text{ at }v% \end{array} \\ 1\text{ or }2 & \begin{array}{l} \text{if }E\text{ has non-split multiplicative reduction} \\ \text{of type }I_{n}\text{ at }v% \end{array}% \end{cases}% $ \medskip If $E$ acquires semi-stable reduction over $L_{z},$ then: \begin{enumerate} \item If $E$ has split multiplicative reduction of type $I_{n}$ over $K_{v},$ then:\newline $c(E/\left( L^{H}\right) _{w})=n.e_{H}.$ \item If $E$ has non-split multiplicative reduction of type $I_{n}$ over $% K_{v},$ then:\newline $c(E/\left( L^{H}\right) _{w})=% \begin{cases} n.e_{H} & \begin{tabular}{l} if $E$ has split multiplicative reduction \\ $\text{over }\left( L^{H}\right) _{w}$% \end{tabular} \\ 1\text{ or }2 & \text{otherwise.}% \end{cases}% $ \item If $E$ has potentially good reduction, then $c(E/\left( L^{H}\right) _{w})=1,2,3$ or $4.$ \item If $E$ has additive and potentially multiplicative reduction then:% \newline $c(E/\left( L^{H}\right) _{w})=% \begin{cases} n.e_{H} & \begin{array}{l} \text{if }E\text{ has split multiplicative reduction} \\ \text{of type }I_{n}\text{ over }\left( L^{H}\right) _{w}\text{ and }% l_{v}\neq 2\text{.}% \end{array} \\ 1,2,3\text{ or }4 & \begin{array}{c} \text{otherwise.}% \end{array}% \end{cases}% $ \end{enumerate} \bigskip The following few remarks will be used in the subsequent computations. \begin{remark} \label{rem1}If $w_{1}$ and $w_{2}$ are two places of $L$ above the same $v,$ then:\newline $c_{w_{1}}(E/L)=c_{w_{2}}(E/L).$\newline In particular:% \begin{equation*} \begin{cases} C_{v}(\{1\})=C_{w}(E/L)^{r} \\ C_{v}(C_{p})=C_{w^{\prime }}(E/L^{C_{p}})^{r^{\prime }}\text{,}% \end{cases}% \end{equation*}% where $r=$the number of places $w$ of $L$ such that $w\mid v$ and $r^{\prime }=$the number of places $w^{\prime }$ of $L^{C_{p}}$ such that $w^{\prime }\mid v$. \end{remark} \begin{remark} \label{rem2}If $E/K$ has potentially good reduction at $v,$ then:\newline $\forall w$ (resp. $w^{\prime }$) place of $L$ (of $L^{C_{p}}$), $c_{w}(E/L)$ ($c_{w^{\prime }}(E/L^{C_{p}})$) $\in \left\{ 1,..,4\right\} $,\newline and therefore $ord_{p}\left( c_{v}\right) =0$ and $(-1)^{ord_{p}\left( C_{v}\right) }=(-1)^{ord_{p}\left( \frac{\omega (\left\{ 1\right\} )}{\omega (C_{p})}\right) }.$ \end{remark} \begin{remark} \label{rem3}If the reduction of $E/K$ at $v$ is semi-stable, then $\forall H$ subgroup of $D_{2p},$ $\delta _{H}=\delta .e_{H}$ and therefore $\omega (H)=1 $ and $(-1)^{ord_{p}\left( C_{v}\right) }=(-1)^{ord_{p}\left( c_{v}\right) }. $ \end{remark} \begin{remark} \label{rem4}If $v\nmid p$ (i.e. $p\neq l_{v},$ $p$ is fixed, $l_{v}$ is variable), then $ord_{p}\left( \omega (H)\right) =0$ and $% (-1)^{ord_{p}\left( C_{v}\right) }=(-1)^{ord_{p}\left( c_{v}\right) }$. \end{remark} \begin{remark} \label{rem5}By the previous two remarks, if $E/K$ has good reduction at $v,$ then: $(-1)^{ord_{p}\left( C_{v}\right) }=1.$\newline As $\frac{W(E/K_{v},\tau _{v})}{W(E/K_{v},(1\oplus \eta )_{v})}=\frac{\det \tau _{v}(-1)}{\det (1\oplus \eta )_{v}(-1)}=1$ in the case of good reduction,$\smallskip $\newline we have the desired equality $(2)$ in the case of good reduction at $v.$ \end{remark} \begin{remark} \label{rem6}From \ref{rem2}\ and \ref{rem4} we deduce that the only case that needs the calculation of both $\omega (H)$ and $c_{w}(E/L^{H})$ is the case of additive potentially multiplicative reduction at $v\mid p.$ \end{remark} \medskip \subsubsection{The cases $G_{v}=\{1\}$ and $G_{v}=C_{p}\label{cases 1 et C_p} $} In these cases, $C_{v}(\{1\})$ and $C_{v}(C_{p})$ are squares, so $% ord_{p}\left( C_{v}\right) \equiv 0\ \left( \func{mod}2\right) .$\newline $\bullet $ If $G_{v}=\{1\}$, $res_{Gal(L_{z}/K_{v})}\tau =1\oplus 1=(1\oplus \eta )_{v}$, hence $\frac{W(E/K_{v},\tau _{v})}{W(E/K_{v},(1\oplus \eta )_{v})}=1.$\newline $\bullet $ If $G_{v}=C_{p}$, $(1\oplus \eta )_{v}=1\oplus 1$ and $\tau _{v}=\chi \oplus \chi ^{\ast }$, so% \begin{equation*} W(E/K_{v},\tau _{v})=1=W(E/K_{v},(1\oplus \eta )_{v})\text{ (see \cite{Dok2} lemma A.1 p.47)}. \end{equation*}% As a result, in both cases we have: $\frac{W(E/K_{v},\tau _{v})}{% W(E/K_{v},(1\oplus \eta )_{v})}=1=(-1)^{ord_{p}\left( C_{v}\right) }.$ \subsubsection{The case $G_{v}=D_{2}\label{case D_2}$} We have $\tau _{v}=(1\oplus \eta )_{v}$, so $\frac{W(E/K_{v},\tau _{v})}{% W(E/K_{v},(1\oplus \eta )_{v})}=1.$ On the other hand, in this case,\newline $\forall w^{\prime }\mid v$ place of $L^{C_{p}}$ and $\forall w\mid w^{\prime }$ place of $L$, \begin{tabular}{c} $L_{w}$ \\ $\shortparallel $ \\ $\left( L^{C_{p}}\right) _{w^{\prime }}$ \\ $\hspace{0.08in}\mid {2}$ \\ $K_{v}$% \end{tabular} In particular, $C_{v}(\{1\})=C_{v}(C_{p})^{p},$ therefore $% C_{v}=C_{v}(C_{p})^{p-1}$ and\newline $ord_{p}\left( C_{v}\right) =0.$ Finally, we get: $\frac{W(E/K_{v},\tau _{v})}{W(E/K_{v},(1\oplus \eta )_{v})}% =1=(-1)^{ord_{p}\left( C_{v}\right) }.$ \subsubsection{The case $G_{v}=D_{2p}\label{case D_2p}$} Denote by $w$ (resp $z$) the unique place of $L^{C_{p}}$ (resp $L$) above $v$% .\newline In this case, there are two possibilities for the inertia group of $G_{v},$ $% I_{v}=C_{p}$ or $D_{2p}$ (because $I_{v}$ is a normal subgroup of $% G_{v}=D_{2p}$ and $G_{v}/I_{v}$ is cyclic).\newline Furthermore, if $l_{v}\neq p$ then $I_{v}=C_{p}:$ - For $l_{v}\neq 2$ because the inertia group of a tamely ramified extension is cyclic. - For $l_{v}=2$ because the case $I_{v}=D_{2p}$, $I_{v}^{wild}=D_{2}$ (the wild inertia group) is impossible since $I_{v}^{wild}$ is normal in $I_{v}$. \paragraph{2.3.4.1\hspace{0.2cm}Computation of $(-1)^{ord_{p}\left( C_{v}\right) }\label{computation ordC_v}$} \begin{enumerate} \item \label{cas1 ordC_v}If $E/K_{v}$ has potentially multiplicative reduction: \begin{enumerate} \item If $E/K_{v}$ acquires split multiplicative reduction of type $I_{n}$ over $L_{z}$ (and therefore over $\left( L^{C_{p}}\right) _{w}$), then:% \newline $C_{v}(\{1\})=c_{w}(E/L_{z})% \begin{tabular}[t]{l} $=e_{L_{z}/\left( L^{C_{p}}\right) _{w}}\times c_{w^{\prime }}(E/\left( L^{C_{p}}\right) _{w})$ \\ $=\dfrac{e_{\{1\}}}{e_{C_{p}}}\times c_{v}(E/K)C_{v}(C_{p})$ \\ $=\dfrac{e_{\{1\}}}{e_{C_{p}}}\times C_{v}(C_{p})$% \end{tabular}% $\newline but $% \begin{cases} \text{if }I_{v}=C_{p}\text{ then }e_{\{1\}}=p\text{ and }e_{C_{p}}=1 \\ \text{if }I_{v}=D_{2p}\text{ then }e_{\{1\}}=2p\text{ and }e_{C_{p}}=2% \end{cases}% $\newline In both cases we get: $C_{v}=p$ and $(-1)^{ord_{p}\left( C_{v}\right) }=-1.$ \item If $E/K_{v}$ does not acquire split multiplicative reduction of type $% I_{n}$ over $L_{z}$ (and therefore nor over $\left( L^{C_{p}}\right) _{w}$), then:\newline $c_{v}(\{1\})$, $c_{v}(C_{p})$ $\in \left\{ 1,2,3,4\right\} $ and $% ord_{p}\left( \dfrac{\omega \left( \left\{ 1\right\} \right) }{\omega \left( C_{p}\right) }\right) \equiv 0$ $\left( \func{mod}2\right) .\smallskip $% \newline The second claim is a consequence of Remark \ref{rem4} in the case $% l_{v}\neq p$.\newline In the case $l_{v}=p,$ we have to distinguish two cases: \begin{enumerate} \item If $E/K_{v}$ acquires non-split multiplicative reduction of type $% I_{n} $ over $L_{z}$ (and therefore over $\left( L^{C_{p}}\right) _{w}$), then $\delta _{\left\{ 1\right\} }=\delta _{C_{p}}.$\newline Furthermore, $f_{C_{p}}=f_{\left\{ 1\right\} }=1$ or $2$ and $\dfrac{\omega \left( \left\{ 1\right\} \right) }{\omega \left( C_{p}\right) }=q^{\delta f(e_{\left\{ 1\right\} }-e_{C_{p}})},$ so $ord_{p}\left( \dfrac{\omega \left( \left\{ 1\right\} \right) }{\omega \left( C_{p}\right) }\right) \equiv 0$ $\left( \func{mod}2\right) $ (because $p-1\mid \left( e_{\left\{ 1\right\} }-e_{C_{p}}\right) $). \item If $E/K_{v},E/\left( L^{C_{p}}\right) _{w}$ and $E/L_{z}$ have additive reduction (of type $I_{n}^{\ast }$):\newline $\bullet $ If $I_{v}=C_{p}$, then $f_{C_{p}}=f_{\left\{ 1\right\} }=2$ and the result follows.\newline $\bullet $ if $I_{v}=D_{2p}$, since $p\geq 5,$ $E$ becomes of type $% I_{2n}^{\ast }$ over $\left( L^{C_{p}}\right) _{w}$ and $I_{2pn}^{\ast }$ over $L_{z}$ and we get: $ord_{p}\left( \omega \left( \left\{ 1\right\} \right) \right) =ord_{p}\left( \omega \left( C_{p}\right) \right) \equiv 0$ $% \left( \func{mod}2\right) .$ \end{enumerate} \end{enumerate} \textit{To sum up, in the case of potentially multiplicative reduction:}% \newline $(-1)^{ord_{p}\left( C_{v}\right) }=% \begin{cases} -1 & \text{if }E/(L^{C_{p}})\text{ has split multiplicative reduction} \\ 1 & \text{otherwise.}% \end{cases}% $ \item If $E/K_{v}$ has potentially good reduction, then: \begin{enumerate} \item If $I_{v}=C_{p}$ (i.e. $e_{\{1\}}=p$ and $e_{C_{p}}=1):$\newline We get: $f_{\{1\}}=f_{C_{p}}=2$ so $ord_{p}(\omega (C_{p}))\equiv ord_{p}(\omega (\{1\}))\equiv 0$ $\left( \func{mod}2\right) \smallskip $% \newline and therefore $(-1)^{ord_{p}\left( C_{v}\right) }=1$ (see Remark \ref{rem2}). \item If $I_{v}=D_{2p}$ (i.e. $e_{\{1\}}=2p$, $e_{C_{p}}=2$ and $l_{v}=p):$% \newline We get: $\frac{C_{v}(\{1\})}{C_{v}(C_{p})}=\frac{\omega (\{1\})}{\omega (C_{p})}=q_{v}^{\left\lfloor \frac{\delta .e_{\left\{ 1\right\} }}{12}% \right\rfloor -\left\lfloor \frac{\delta .e_{C_{p}}}{12}\right\rfloor }=q_{v}^{\left\lfloor \frac{\delta .2p}{12}\right\rfloor -\left\lfloor \frac{% \delta .2}{12}\right\rfloor }.$ \begin{enumerate} \item If $q_{v}$ is an even power of $p,$ then\newline $(-1)^{ord_{p}\left( C_{v}\right) }=(-1)^{ord_{p}\left( \frac{\omega (\{1\})% }{\omega (C_{p})}\right) }=1.$ \item \label{tableau}If $q_{v}$ is an odd power of $p:$\newline A computation of $\left\lfloor \frac{\delta .2p}{12}\right\rfloor $ and $% \left\lfloor \frac{\delta .2}{12}\right\rfloor $ depending on $p$ modulo $12$ gives the following table:\smallskip \newline Table of values of $(-1)^{ord_{p}\left( C_{v}\right) }$ depending on the Kodaira symbol of the curve (and the value of $\mathfrak{e}=\frac{12}{% \limfunc{pgcd}(\delta ,12)}$) and $p$ $\func{mod}12$:% \begin{equation*} \begin{tabular}{|l|c|c|c|c|} \hline $p$ $\func{mod}12$ & $1$ & $5$ & $7$ & $11$ \\ \hline $II,II^{\ast }$ ($\mathfrak{e}=6$) & 1 & -1 & 1 & -1 \\ \hline $III,III^{\ast }$ ($\mathfrak{e}=4$) & 1 & 1 & -1 & -1 \\ \hline $IV,IV^{\ast }$ ($\mathfrak{e}=3$) & 1 & -1 & 1 & -1 \\ \hline $I_{o}^{\ast }$ ($\mathfrak{e}=2$) & 1 & 1 & 1 & 1 \\ \hline \end{tabular}% \end{equation*}% In relation to the above table it may be useful to recall the following fact: if the residue characteristic of $K_{v}$ is $>3,$ then we have the following correspondence between $\mathfrak{e}=\frac{12}{\limfunc{pgcd}% (\delta ,12)},$ the valuation of the minimal discriminant $\delta $\ and the Kodaira symbols:% \begin{equation*} \begin{tabular}[t]{lll} $\mathfrak{e}=1$ & $\Leftrightarrow \delta =0$ & $\Leftrightarrow E$ is of type $I_{0}$ \\ $\mathfrak{e}=2$ & $\Leftrightarrow \delta =6$ & $\Leftrightarrow E$ is of type $I_{0}^{\ast }$ \\ $\mathfrak{e}=3$ & $\Leftrightarrow \delta =4$ or $8$ & $\Leftrightarrow E$ is of type $IV$ or $IV^{\ast }$ \\ $\mathfrak{e}=4$ & $\Leftrightarrow \delta =3$ or $9$ & $\Leftrightarrow E$ is of type $III$ or $III^{\ast }$ \\ $\mathfrak{e}=6$ & $\Leftrightarrow \delta =2$ or $10$ & $\Leftrightarrow E$ is of type $II$ or $II^{\ast }$.% \end{tabular}% \end{equation*} \end{enumerate} \end{enumerate} \end{enumerate} \paragraph{2.3.4.2\hspace{0.2cm}Computation of $\frac{W(E/K_{v},\protect\tau % _{v})}{W(E/K_{v},\left( 1\oplus \protect\eta \right) _{v})}$} \begin{enumerate} \item \label{Cas1}The case of potentially multiplicative reduction:\newline We have an explicit formula of Rohrlich (see \cite{Roh2} Th.2 (ii) p.329):% \begin{equation*} W(E/K_{v},\sigma )=\det \sigma (-1)\chi (-1)^{\dim \sigma }(-1)^{\left\langle \chi ,\sigma \right\rangle }\text{,} \end{equation*}% where $\chi $ is the character of $K_{v}^{\ast }$ associated to the extension $K_{v}(\sqrt{-c_{6}})$ of $K_{v}$ ($c_{6}$ is the classical factor, see \cite{Silv1} p.46).\newline Since $\dim \tau _{v}=\dim 1\oplus \eta =2$, $\det (\tau _{v})=\det (1\oplus \eta )$ and $\left\langle \chi ,\tau _{v}\right\rangle =0$, we get:% \begin{equation*} \frac{W(E/K_{v},\tau _{v})}{W(E/K_{v},\left( 1\oplus \eta \right) _{v})}=% \frac{(-1)^{\left\langle \chi ,\tau _{v}\right\rangle }}{(-1)^{\left\langle \chi ,\left( 1\oplus \eta \right) _{v}\right\rangle }}=\frac{1}{% (-1)^{\left\langle \chi ,\left( 1\oplus \eta \right) _{v}\right\rangle }}% =(-1)^{\left\langle \chi ,\left( 1\oplus \eta \right) _{v}\right\rangle }% \text{.} \end{equation*}% \medskip \begin{enumerate} \item If the reduction of $E/K_{v}$ is split multiplicative (i.e. $\chi =1$):% \newline Then $(-1)^{\left\langle \chi ,\left( 1\oplus \eta \right) _{v}\right\rangle }=-1$. \item If the reduction of $E/K_{v}$ is non-split multiplicative (i.e. $\chi $ is an unramified quadratic character): \begin{enumerate} \item If $E$ acquires split multiplicative reduction over $L_{z}$ (and therefore over $\left( L^{C_{p}}\right) _{w}$), then $\eta _{v}=\chi $, hence $(-1)^{\left\langle \chi ,\left( 1\oplus \eta \right) _{v}\right\rangle }=-1.$ \item If $E$ acquires non-split multiplicative reduction over $L_{z}$ (and therefore over $\left( L^{C_{p}}\right) _{w}$), then $\eta _{v}\neq \chi $, hence $(-1)^{\left\langle \chi ,\left( 1\oplus \eta \right) _{v}\right\rangle }=1.$ \end{enumerate} \item If the reduction of $E/K_{v}$ is additive (i.e. $\chi $ is a ramified quadratic character) \begin{enumerate} \item If $E$ acquires split multiplicative reduction over $L_{z}$ (and therefore over $\left( L^{C_{p}}\right) _{w}$), then $\eta _{v}=\chi $, hence $(-1)^{\left\langle \chi ,\left( 1\oplus \eta \right) _{v}\right\rangle }=-1.$ \item If $E$ acquires non-split multiplicative reduction over $L_{z}$ (and therefore over $\left( L^{C_{p}}\right) _{w}$), then $\eta _{v}\neq \chi $, hence $(-1)^{\left\langle \chi ,\left( 1\oplus \eta \right) _{v}\right\rangle }=1.$ \end{enumerate} \end{enumerate} \textit{To sum up, in the case of potentially multiplicative reduction:}% \newline $\frac{W(E/K_{v},\tau _{v})}{W(E/K_{v},\left( 1\oplus \eta \right) _{v})}$% \begin{tabular}[t]{l} $=% \begin{cases} -1 & \text{if }E/(L^{C_{p}})\text{\ has split multiplicative reduction} \\ 1 & \text{otherwise.}% \end{cases}% $ \\ $=(-1)^{ord_{p}\left( C_{v}\right) }$, by \ref{computation ordC_v}.1.\ref% {cas1 ordC_v}% \end{tabular} \item The case of potentially good reduction:\newline Here we have to distinguish the cases $l_{v}=p$ and $l_{v}\neq p.$ \begin{enumerate} \item The case $l_{v}=p.$\newline We have again an explicit formula of Rohrlich, since $p\geq 5$ (see \cite% {Roh2}, Th.2 (iii) p.329):\newline We use the following notation:{}\newline $\bullet $ $q=p^{r}$ the cardinality of the residue field residue degree of $% K_{v}$\newline $\bullet $ $\mathfrak{e}=\frac{12}{\limfunc{pgcd}(\delta ,12)}$\newline $\bullet $ $\epsilon =% \begin{cases} 1 & \text{if }r\text{ is even or }\mathfrak{e}=1 \\ \left( \frac{-1}{p}\right) & \text{if }r\text{ is odd and }\mathfrak{e}=2% \text{ or }6 \\ \left( \frac{-3}{p}\right) & \text{if }r\text{ is odd and }\mathfrak{e}=3 \\ \left( \frac{-2}{p}\right) & \text{if }r\text{ is odd and }\mathfrak{e}=4% \text{.}% \end{cases}% $\newline Then $\forall \sigma $ a self-dual representation of $Gal(\overline{K}% _{v}/K_{v})$ with finite image:% \begin{equation*} W(E/K_{v},\sigma )=% \begin{cases} \alpha (\sigma ,\epsilon ) & \begin{tabular}{l} if $q\equiv 1[\mathfrak{e}]$% \end{tabular} \\ \alpha (\sigma ,\epsilon )(-1)^{\left\langle 1+\eta _{nr}+\hat{\sigma}% _{e},\sigma \right\rangle } & \begin{array}{l} \text{if }q\equiv -1[\mathfrak{e}] \\ \text{and }\mathfrak{e}=3,4,6,% \end{array}% \end{cases}% \end{equation*}% where $\eta _{nr}$ is the unramified quadratic character, $\hat{\sigma}_{e}$ is an irreductible representation of degree 2 of $D_{2\mathfrak{e}}$ and $% \alpha (\sigma ,\epsilon ):=(\det \sigma )(-1)\epsilon ^{\dim \sigma }$.% \newline Since $\dim \tau _{v}=\dim \left( 1\oplus \eta \right) _{v}=2$ and $\det \tau _{v}=\det \left( 1\oplus \eta \right) _{v}$,\newline $\alpha (\left( 1\oplus \eta \right) _{v},\epsilon )=\alpha (\tau _{v},\epsilon )$ and we get:\newline \begin{tabular}{ll} $\frac{W(E/K_{v},\tau _{v})}{W(E/K_{v},\left( 1\oplus \eta \right) _{v})}$ & $=% \begin{cases} 1 & \begin{tabular}{l} if $q\equiv 1[\mathfrak{e}]$% \end{tabular} \\ (-1)^{\left\langle 1+\eta _{nr}+\hat{\sigma}_{e},1+\eta _{v}+\tau _{v}\right\rangle } & \begin{array}{l} \text{if }q\equiv -1[\mathfrak{e}] \\ \text{and }\mathfrak{e}=3,4,6,% \end{array}% \end{cases}% \smallskip $ \\ & $=% \begin{cases} 1 & \begin{tabular}{l} if $q\equiv 1[\mathfrak{e}]$% \end{tabular} \\ (-1)^{\left\langle 1+\eta _{nr},1+\eta _{v}\right\rangle } & \begin{array}{l} \text{if }q\equiv -1[\mathfrak{e}] \\ \text{and }\mathfrak{e}=3,4,6,% \end{array}% \end{cases}% $ \\ & ($\left\langle \hat{\sigma}_{e},\tau _{v}\right\rangle =0$ since $% \mathfrak{e}=3,4,6$ and $p\geq 5$).$\bigskip $% \end{tabular} \begin{enumerate} \item If $r$ is even, then $q\equiv 1[\mathfrak{e}]$ $\forall \mathfrak{e}% \in \left\{ 2,3,4,6\right\} $ and therefore% \begin{equation*} \frac{W(E/K_{v},\tau _{v})}{W(E/K_{v},\left( 1\oplus \eta \right) _{v})}% =1=(-1)^{ord_{p}\left( C_{v}\right) }, \end{equation*}% by 2.b.i (in section 2.3.4.1). \item If $r$ is odd, then $q\equiv 1[\mathfrak{e}]\Longleftrightarrow p\equiv 1[\mathfrak{e}]$ and:% \begin{equation*} \frac{W(E/K_{v},\tau _{v})}{W(E/K_{v},\left( 1\oplus \eta \right) _{v})}=% \begin{cases} 1 & \begin{tabular}{l} if $q\equiv 1[\mathfrak{e}]$% \end{tabular} \\ (-1)^{\left\langle 1+\eta _{nr},1+\eta _{v}\right\rangle } & \begin{array}{l} \text{if }q\equiv -1[\mathfrak{e}]\text{ and} \\ \mathfrak{e}=3,4,6\text{.}% \end{array}% \end{cases}% \end{equation*}% \bigskip \begin{enumerate} \item If $I_{v}=C_{p},$ then $\eta _{nr}=\eta _{v}$ and $\frac{% W(E/K_{v},\tau _{v})}{W(E/K_{v},\left( 1\oplus \eta \right) _{v})}=1$. \item If $I_{v}=D_{2p},$ then $\eta _{nr}\neq \eta _{v}$ and:% \begin{equation*} \frac{W(E/K_{v},\tau _{v})}{W(E/K_{v},\left( 1\oplus \eta \right) _{v})}=% \begin{cases} 1 & \text{if }q\equiv 1[\mathfrak{e}] \\ -1 & \text{if }q\equiv -1[\mathfrak{e}]\text{ and }\mathfrak{e}=3,4,6\text{.}% \end{cases}% \end{equation*} \end{enumerate} In both cases, we obtain for the values of $\frac{W(E/K_{v},\tau _{v})}{% W(E/K_{v},\left( 1\oplus \eta \right) _{v})}$ exactly the same table as for the values of $(-1)^{ord_{p}\left( C_{v}\right) },$ depending on $p$ modulo $% 12$:\medskip \newline Table of values of $\frac{W(E/K_{v},\tau _{v})}{W(E/K_{v},\left( 1\oplus \eta \right) _{v})}$ depending on the Kodaira symbol of the curve (and the value of $\mathfrak{e}=\frac{12}{\limfunc{pgcd}(\delta ,12)}$) and $p$ $% \func{mod}12$:% \begin{equation*} \begin{tabular}{|l|l|l|l|l|} \hline $p$ $\func{mod}12$ & $1$ & $5$ & $7$ & $11$ \\ \hline $II,II^{\ast }$ ($\mathfrak{e}=6$) & \multicolumn{1}{|c|}{1} & \multicolumn{1}{|c|}{-1} & \multicolumn{1}{|c|}{1} & \multicolumn{1}{|c|}{-1} \\ \hline $III,III^{\ast }$ ($\mathfrak{e}=4$) & \multicolumn{1}{|c|}{1} & \multicolumn{1}{|c|}{1} & \multicolumn{1}{|c|}{-1} & \multicolumn{1}{|c|}{-1} \\ \hline $IV,IV^{\ast }$ ($\mathfrak{e}=3$) & \multicolumn{1}{|c|}{1} & \multicolumn{1}{|c|}{-1} & \multicolumn{1}{|c|}{1} & \multicolumn{1}{|c|}{-1} \\ \hline $I_{o}^{\ast }$ ($\mathfrak{e}=2$) & \multicolumn{1}{|c|}{1} & \multicolumn{1}{|c|}{1} & \multicolumn{1}{|c|}{1} & \multicolumn{1}{|c|}{1} \\ \hline \end{tabular}% \end{equation*} \end{enumerate} \item The case $l_{v}\neq p:$\newline In this case, the explicit formula of Rohrlich cannot be used, since $l_{v}$ can be $2$ or $3.$\newline Let $\sigma $ be a representation $\sigma $ $:$ $Gal(\overline{K}% _{v}/K_{v})\rightarrow GL(V_{\sigma })$ with finite image; let $\sigma _{E/K_{v}}^{\prime }:WD(\overline{K}_{v}/K_{v})\rightarrow GL(V)$ be the representation of the Weil-Deligne group associated to the elliptic curve given by $\left( \sigma _{E/K_{v}},N\right) =\left( \sigma _{E/K_{v}},0\right) $ (because the reduction is potentially good), therefore this is simply a representation of the Weil group $W(\bar{K}_{v}/K_{v})$ (because $N=0$) and \begin{equation*} \sigma _{E/K_{v}}^{\prime }\otimes \sigma =\sigma _{E/K_{v}}\otimes \sigma :W(\overline{K}_{v}/K_{v})\rightarrow GL(W), \end{equation*}% where $W=V\otimes V_{\sigma },$ is also a representation of the Weil group.% \newline We first recall the link between $\varepsilon $-factors and root numbers:% \begin{equation*} W(E/K_{v},\sigma )=\dfrac{\varepsilon (\sigma _{E/K_{v}}\otimes \sigma ,\psi ,dx)}{\left\vert \varepsilon (\sigma _{E/K_{v}}\otimes \sigma ,\psi ,dx)\right\vert }=\varepsilon (\sigma _{E/K_{v}}^{\prime }\otimes \sigma ,\psi ,dx_{\psi }), \end{equation*}% where $dx$ is any Haar measure, $\psi $ is any additive character of $K_{v}$ and $dx_{\psi }$ the self-dual Haar measure with respect to $\psi $ on $% K_{v} $.\newline Here, we choose an additive character $\psi $ for which the Haar measure $% dx_{\psi }$ takes values (on open compact subsets of $K_{v}$) in $\mathbb{Z}% _{p}[\zeta _{p}],$ where $\zeta _{p}$ is a primitive $p$-th root of unity. For example, if the conductor of $\psi $ is trivial, then the values of $% dx_{\psi }$ lie in $l_{v}^{\mathbb{Z}}\cup \{0\}\subset \mathbb{Z}_{p}[\zeta _{p}]$.\newline In one of his articles (\cite{Del} p.548), Deligne gives a description of the $\varepsilon $-factors in terms of $\varepsilon _{0}$-factors; in our settings this gives:% \begin{equation*} \varepsilon (\sigma _{E/K_{v}}\otimes \sigma ,\psi ,dx_{\psi })=\varepsilon _{0}(\sigma _{E/K_{v}}\otimes \sigma ,\psi ,dx_{\psi })\det (-\nu (\phi )\mid W^{I(v)}), \end{equation*}% where $\phi $ is the geometric Frobenius at $v$ and $I(v)=Gal(\bar{K}% _{v}/K_{v}^{ur}).$\newline Recall that, since $l_{v}\neq p,$ the inertia group of $D_{2p}$ is $% I_{v}=C_{p}$.\medskip \begin{enumerate} \item If $E$ has additive reduction, denote by $F$ the smallest Galois extension of $K_{v}^{ur}$ such that $E$ has good reduction over $F$ and set $% \Phi =Gal(F/K_{v}^{ur})$; then the restiction of $\sigma _{E/K_{v}}$ to $% I(v) $ factors through $\Phi .$\newline It is known that:\newline $\bullet $ For $l_{v}\geq 5,$ $\Phi $ is cyclic of order $\mathfrak{e}=\frac{% 12}{\limfunc{pgcd}(\delta ,12)}$ (dividing 12).\newline $\bullet $ For $l_{v}=3,$ $\left\vert \Phi \right\vert \in \left\{ 2,3,4,6,12\right\} .$\newline $\bullet $ For $l_{v}=2,$ $\left\vert \Phi \right\vert \in \left\{ 2,3,4,6,8,24\right\} .$\newline For a more precise description of $\Phi $, see, for example, \cite{Bil} or \cite{Kraus}.\newline The representation $\sigma _{E/K}\otimes \sigma $ ($\sigma =\tau _{v}$ or $% \left( 1\oplus \eta \right) _{v}$) restricted to $I(v)$ factors through a quotient $H$ of $I(v)$ which admits $\Phi $ and $C_{p}$ as quotients.\newline We have:% \begin{equation*} \left( V\otimes V_{\sigma }\right) ^{I(v)}=\left( V\otimes V_{\sigma }\right) ^{H}=Hom_{H}(V^{\ast },V_{\sigma })=Hom((V^{\Phi })^{\ast },V_{\sigma }^{C_{p}}) \end{equation*}% because $H$ acts on $V$ (resp. on $V_{\sigma }$) through its quotient $\Phi $ (resp. $C_{p}$) and $\left\vert \Phi \right\vert $ is prime to $p$.\newline Futhermore, $V^{H}=V^{\Phi }=\left\{ 0\right\} $ since $E$ has additive reduction, hence% \begin{equation*} \left( V\otimes V_{\sigma }\right) ^{I(v)}=0,\hspace{0.3cm}\det \left( -\left( \sigma _{E/K_{v}}^{\prime }\otimes \sigma \right) (\phi )\mid \left( V\otimes V_{\sigma }\right) ^{I(v)}\right) =1 \end{equation*}% and% \begin{equation*} W(E/K_{v},\sigma )=\varepsilon _{0}(\sigma _{E/K_{v}}\otimes \sigma ,\psi ,dx_{\psi })\qquad (\sigma =\tau _{v},(1\oplus \eta )_{v}).\leqno(3) \end{equation*}% \bigskip Deligne also gives congruence results for these $\varepsilon _{0}$ (% \cite{Del} p.556-557). Since $\chi \equiv 1$ $\func{mod}(1-\zeta _{p})$, we deduce $I(\chi )\equiv I(1)$ $\func{mod}(1-\zeta _{p})$ and $\sigma _{E/K_{v}}^{\prime }\otimes \tau _{v}\equiv \sigma _{E/K_{v}}^{\prime }\otimes \left( 1\oplus \eta \right) _{v}$ $\func{mod}(1-\zeta _{p})$. So according to Deligne, $\varepsilon _{0}(\sigma _{E/K_{v}}^{\prime }\otimes \tau _{v},\psi ,dx_{\psi })$ and $\varepsilon _{0}(\sigma _{E/K_{v}}^{\prime }\otimes \left( 1\oplus \eta \right) _{v},\psi ,dx_{\psi })$ are two elements of $\{\pm 1\}$ (by (3)), which are congruent modulo $(1-\zeta _{p})$% , hence the are equal. As a result, \begin{equation*} \frac{W(E/K_{v},\tau _{v})}{W(E/K_{v},\left( 1\oplus \eta \right) _{v})}=1. \end{equation*} \item If $E$ has good reduction, then $\sigma _{E/K_{v}}$ is unramified.% \newline Then we have:\newline $\varepsilon (\sigma _{E/K_{v}}\otimes \tau _{v},\psi ,dx)=\varepsilon (\tau _{v},\psi ,dx)^{\dim \sigma _{E/K_{v}}}\det \sigma _{E/K_{v}}(\Phi ^{m(\tau _{v},\psi )}),$ where $m(\tau _{v},\psi )\in \mathbb{N}$ depends on conductors of both $\tau _{v}$ and $\psi $, and the dimension of $\tau _{v}$ (see \cite{Tate1} 3.4.6 p.15), therefore:% \begin{equation*} W(E/K_{v},\tau _{v})=W(\sigma _{E/K_{v}}\otimes \tau _{v})=\dfrac{% \varepsilon (\sigma _{E/K_{v}}\otimes \tau _{v},\psi ,dx)}{\left\vert \varepsilon (\sigma _{E/K_{v}}\otimes \tau _{v},\psi ,dx)\right\vert }=1% \text{,} \end{equation*}% since $\det \sigma _{E/K_{v}}=1$, $W(\tau _{v})=\dfrac{\varepsilon (\tau _{v},\psi ,dx)}{\left\vert \varepsilon (\tau _{v},\psi ,dx)\right\vert }=\pm 1$ (because $\det \tau _{v}=1,$ see Proposition p.145 \cite{Roh1}) and $\dim \sigma _{E/K_{v}}=2.$\newline Similarly, $W(E/K_{v},\left( 1\oplus \eta \right) _{v})=1,$ so $\frac{% W(E/K_{v},\tau _{v})}{W(E/K_{v},\left( 1\oplus \eta \right) _{v})}=1.$ \end{enumerate} In both cases i) and ii) we also have $(-1)^{ord_{p}(C_{v})}=1$ by 2.a. (in section 2.3.4.1). \end{enumerate} \end{enumerate} \bigskip To sum up, we have, for each finite prime $v$ of $K$,% \begin{equation*} \frac{W(E/K_{v},\tau _{v})}{W(E/K_{v},\left( 1\oplus \eta \right) _{v})}% =(-1)^{ord_{p}\left( C_{v}\right) }. \end{equation*}% This completes the proof of Theorem \ref{theo}.\hspace{5cm}$\blacksquare $ \begin{remark} This proof can be adjusted to work in the case $Gal(L/K)\simeq D_{2p^{n}},$ the computations are almost the same. The idea to reduce the proof to the case of a $D_{2p}$-extension, using Galois invariance of Rohrlich, was suggested to me by Tim Dokchitser. \end{remark} \section{Appendix} The purpose of this appendix is to make a small improvement on Theorem 6.7 of \cite{Dok3}. The interest of this improvement is that Proposition 6.12 of \cite{Dok3} (which is the same statement as Theorem \ref{theo2} for $p\equiv 3$ $\func{mod}4)$ will no longer rely on the "truly painful case of additive reduction" anymore (see \cite{Dok2} p.30). In fact, we use the passage to the global case to avoid all places of additive reduction, not just those above $2$ and $3$. Since we have proved the result for $p\geq 5$ (Theorem % \ref{theo2}) without using any global parity results at all, for us this is of interest essentially in the case $p=3.$\medskip \medskip We start by recalling the definition of an elliptic curve being \textit{% close to} another one: \begin{proposition} \label{Prop courbe proche}Let $\mathcal{E}:$ $% y^{2}+a_{1}xy+a_{3}y=x^{3}+a_{2}x^{2}+a_{4}x+a_{6}$ be an elliptic curve over a non archimedean local field $\mathcal{K}$ (with valuation $v$ and residue characteristic $p$) and $\mathcal{F}/\mathcal{K}$ a finite Galois extension.\newline There exists $\varepsilon >0$ such that every elliptic curve $\mathcal{E}% ^{\prime }:$ $y^{2}+a_{1}^{\prime }xy+a_{3}^{\prime }y=x^{3}+a_{2}^{\prime }x^{2}+a_{4}^{\prime }x+a_{6}^{\prime }$ over $\mathcal{K}$ satisfying:% \begin{equation*} \forall i\left\vert a_{i}^{\prime }-a_{i}\right\vert _{v}<\varepsilon , \end{equation*}% the curves $\mathcal{E}$ and $\mathcal{E}^{\prime }$ have the same conductor, valuation of the minimal discriminant, local Tamagawa factors, $% C(E/\mathcal{F}^{\prime },\frac{dx}{2y+a_{1}x+a_{3}}),$ root numbers and the Tate module as a $Gal(\mathcal{\bar{K}}/\mathcal{K)}$-module for $l\neq p$, in all intermediate fields $\mathcal{F}^{\prime }$ of $\mathcal{F}/\mathcal{K% }$.\newline We will say that $\mathcal{E}^{\prime }$ is \textit{close to} $\mathcal{E}/% \mathcal{K}$. \end{proposition} \begin{proof} This is Proposition 3.3 of \cite{Dok3}.\medskip \end{proof} We now state the minor improvement of Theorem 6.7 of \cite{Dok3}: \begin{theorem} Let $\mathcal{K}$ a local non archimedean field of characteristic $0$ and $% \mathcal{F}/\mathcal{K}$ a finite Galois extension. Let $F/K$ be a Galois extension of totally real fields and $v_{0}$ a place of $K$ such that:% \newline $\bullet $ $v_{0}$ admits a unique place $\bar{v}_{0}$ of $F$ above it% \newline $\bullet $ $K_{v_{0}}\simeq \mathcal{K}$ and $F_{\bar{v}_{0}}\simeq \mathcal{% F}$.\newline Such an extension exists (see Lemma 3.1 of \cite{Dok3}).\newline Let $\mathcal{E}/\mathcal{K}$ be an elliptic curve with additive reduction.% \newline Then there exists an elliptic curve $E/K$ such that:\newline $\bullet $ $E$ has semi-stable reduction for all $w\neq v_{0}$\newline $\bullet $ $j(E)$ is not an integer (i.e. $j(E)\notin \mathcal{O}_{K})$% \newline $\bullet $ $E/K_{v_{0}}$ is \textit{close to} $\mathcal{E}/\mathcal{K}$. \end{theorem} \begin{proof} We first choose an elliptic curve $E/K$ such that $E/K_{v_{0}}$ is \textit{% close to} $\mathcal{E}/\mathcal{K}$ (this is possible, by Proposition \ref% {Prop courbe proche}).\newline Now the goal is to remove all places of additive reduction by changing $E/K$ to an elliptic curve satisfying the three conditions of the theorem.\newline Let $E:y^{2}+a_{1}xy+a_{3}y=x^{3}+a_{2}x^{2}+a_{4}x+a_{6}$ with $a_{i}\in \mathcal{O}_{K}$.\newline If we want a place not to be of additive reduction we have to impose one of the two following conditions:\newline $\bullet $ The valuation $w(\Delta )$ is zero (in this case $w$ is of good reduction).\newline $\bullet $ The valuation $w(c_{4})$ is zero (in this case $w$ is of good or multiplicative reduction depending on $w(\Delta )=0$ or $>0$).\newline \smallskip Let $v\neq v_{0}$ be a place of $K$ not above $2$.\newline To get the condition "$j(E)$ is not an integer" it is sufficient to make $v$ a multiplicative place ($v$ is multiplicative $\Leftrightarrow $ $v(j(E))<0$% ).\newline \textbf{Step 1}:\textbf{\ Make semi-stable all places }$w\neq v_{0}$ \textbf{% above }$2$\newline Denote by $v_{2,1},...,v_{2,r}$ these places.\newline In this case: $\left[ v_{2,i}(a_{1})=0\Rightarrow v_{2,i}(c_{4})=0\text{ (}% c_{4}=(a_{1}^{2}+4a_{2})^{2}-24a_{1}a_{3}-48a_{4}\text{)}\right] .$\newline Let $\mathfrak{p}_{0}$ and $\mathfrak{p}_{2,i}$ be the primes ideals associated to $v_{0}$ and $v_{2,i}.$\newline By the Chinese remainder theorem, there exists $d_{1}\in \mathcal{O}_{K}$ such that:\newline $\bullet $ $d_{1}\equiv 0$ mod $\mathfrak{p}_{0}^{n}$ (i.e. $% v_{0}(d_{1})\geq n).$\newline $\bullet $ $d_{1}\equiv 1-a_{1}$ mod $\mathfrak{p}_{2,i}$ $\forall i\in \left\{ 1,..,r\right\} $ (i.e. $v_{2,i}(a_{1}+d_{1})=0$).\newline $\bullet $ $d_{1}\equiv -a_{1}$ mod $\mathfrak{p}$ ($\mathfrak{p}$ associated to $v\neq v_{0}$).\newline So, if we let $a_{1}^{\prime }=a_{1}+d_{1}$ for $n$ big enough we get the curve $y^{2}+a_{1}^{\prime }xy+a_{3}y=x^{3}+a_{2}x^{2}+a_{4}x+a_{6}$ which is \textit{close to} $\mathcal{E}/\mathcal{K}$, $v_{2,i}(a_{1}^{\prime })=v_{2,i}(a_{1}+d_{1})=0$ $\forall i\in \left\{ 1,..,r\right\} $ and $% v(a_{1}^{\prime })>0.$\newline \textbf{Step 2}: \textbf{Make }$v$\textbf{\ semi-stable}\newline By the Chinese remainder theorem, there exist $d_{2},d_{3},d_{4}\in \mathcal{% O}_{K}$ such that:\newline $\bullet $ \begin{tabular}[t]{l} $d_{2}\equiv 0$ mod $\mathfrak{p}_{0}^{n}$ (i.e. $v_{0}(d_{2})\geq n)$ \\ $d_{2}\equiv 1-a_{2}$ mod $\mathfrak{p}$ (so $v(a_{2}+d_{2})=0$).% \end{tabular}% \newline $\bullet $ \begin{tabular}[t]{l} $d_{3}\equiv 0$ mod $\mathfrak{p}_{0}^{n}$ (i.e. $v_{0}(d_{3})\geq n)$ \\ $d_{3}\equiv -a_{3}$ mod $\mathfrak{p}$ (so $v(a_{3}+d_{3})>0$).% \end{tabular}% \newline $\bullet $ \begin{tabular}[t]{l} $d_{4}\equiv 0$ mod $\mathfrak{p}_{0}^{n}$ (i.e. $v_{0}(d_{4})\geq n)$ \\ $d_{4}\equiv -a_{4}$ mod $\mathfrak{p}$ (so $v(a_{4}+d_{4})>0$).% \end{tabular}% \newline So, if we let $a_{i}^{\prime }=a_{i}+d_{i},$ $i\in \left\{ 2,3,4\right\} $, for $n$ big enough we get:\newline $E^{\prime }:y^{2}+a_{1}^{\prime }xy+a_{3}^{\prime }y=x^{3}+a_{2}^{\prime }x^{2}+a_{4}^{\prime }x+a_{6}$ is \textit{close to} $\mathcal{E}/\mathcal{K}$ (Proposition \ref{Prop courbe proche}).\newline Futhermore :% \begin{tabular}[t]{l} $\bullet $ $c_{4}^{\prime }=(a_{1}^{\prime 2}+4a_{2}^{\prime })^{2}-24a_{1}^{\prime }a_{3}^{\prime }-48a_{4}^{\prime }$ \\ $\bullet $ $v(a_{1}^{\prime })>0$ \\ $\bullet $ $v(a_{3}^{\prime })>0$ \\ $\bullet $ $v(a_{4}^{\prime })>0$ \\ $\bullet $ $v(a_{2}^{\prime })=0$,% \end{tabular}% \newline so $v(c_{4}^{\prime })=0$.\medskip \newline The curve $E^{\prime }:y^{2}+a_{1}^{\prime }xy+a_{3}^{\prime }y=x^{3}+a_{2}^{\prime }x^{2}+a_{4}^{\prime }x+a_{6}$ is \textit{close to} $% \mathcal{E}/\mathcal{K}$, $\forall w\neq v_{0}$ above\textbf{\ }$2$ $% w(c_{4}^{\prime })>0$, and $v(c_{4}^{\prime })=0.$ Since $c_{4}^{\prime }$ does not depend on $a_{6},$ we can modify $a_{6}$ to allow places $w\neq v_{0}$ such that $w(c_{4}^{\prime })>0$ to become places of good reduction (since $c_{4}^{\prime }$ will be unchanged, some places of good reduction can become of multiplicative reduction but not of additive reduction) and such that $v$ is of multiplicative reduction ($v(j(E))<0$).\medskip \newline \textbf{Step 3}. \textbf{Turn additive reduction places into good reduction ones and make }$v$\textbf{\ multiplicative.}\newline Let $v_{1},...,v_{r},v_{r+1},...,v_{t}$ be the places where $% v_{i}(c_{4}^{\prime })>0,$ $v_{i}\neq v_{0}$ ($\neq v$ and not above $2$).% \newline Above, $v_{1},...,v_{r}$ are places of good reduction and $v_{r+1},...,v_{t}$ places of additive reduction of the curve $E^{\prime }$ constructed in step 2.\newline Let $b_{2},$ $b_{4},$ $b_{6},$ $b_{8}$ and $\Delta $ be the following classical quantities associatied to $E^{\prime }$:\newline \begin{tabular}[t]{l} $b_{2}=a_{1}^{\prime 2}+4a_{2}^{\prime }$ \\ $b_{4}=2a_{4}^{\prime }+a_{1}^{\prime }a_{3}^{\prime }$ \\ $b_{6}=a_{3}^{\prime 2}+4a_{6}$ \\ $b_{8}=a_{1}^{\prime 2}a_{6}+4a_{2}^{\prime }a_{6}-a_{1}^{\prime }a_{3}^{\prime }a_{4}^{\prime }+a_{2}^{\prime }a_{3}^{\prime 2}-a_{4}^{\prime 2}$% \end{tabular}% \newline and $\Delta \begin{tabular}[t]{l} $=-b_{2}^{2}b_{8}-8b_{4}^{3}-27b_{6}^{2}+9b_{2}b_{4}b_{6}$ \\ $=\alpha +\beta .a_{6}+16.a_{6}^{2}$, \\ where $\alpha =[-b_{2}^{2}(-a_{1}^{\prime }a_{3}^{\prime }a_{4}^{\prime }+a_{2}^{\prime }a_{3}^{\prime 2}-a_{4}^{\prime 2})-8b_{4}^{3}-27a_{3}^{\prime 4}+9b_{2}b_{4}a_{3}^{\prime 2}]$ \\ and $\beta =[-b_{2}^{3}-216a_{3}^{\prime 2}+36b_{2}b_{4}]$% \end{tabular}% $\newline Let $\gamma =\beta +32.a_{6}$; we know that $16$ is invertible mod $% \mathfrak{p}_{i}$ $\forall i\in \left\{ 1,..,t\right\} $ (because $\mathfrak{% p}_{i}$ is not above 2).\smallskip \newline By the Chinese remainder theorem, there exists $c$ such that:\newline $\bullet $ $c\equiv 0$ mod $\mathfrak{p}_{0}^{n}$ (i.e. $v_{0}(c)\geq n)$% \newline $\bullet $ $c\equiv 0$ mod $\mathfrak{p}_{i}$ $\forall i\in \left\{ 1,..,r\right\} $ (i.e. $v_{i}(c)>0$)\newline $\bullet $ $16c\equiv \alpha _{i}-\gamma $ mod $\mathfrak{p}_{i}$ $\forall i\in \left\{ r+1,..,t\right\} $ (where $\alpha _{i}\neq 0,\gamma $ mod $% \mathfrak{p}_{i})$\newline (i.e. $\forall i\in \left\{ r+1,..,t\right\} ,$ $v_{i}(\gamma +16c)=0$ and $% v_{i}(c)=0$)\newline $\bullet $ $c\equiv -a_{6}$ mod $\mathfrak{p}$ (i.e. $v(a_{6}^{\prime })>0)$.% \newline Finally, if we let $a_{6}^{\prime }=a_{6}+c$ for $n$ big enough, we get:% \newline $E^{\prime \prime }:y^{2}+a_{1}^{\prime }xy+a_{3}^{\prime }y=x^{3}+a_{2}^{\prime }x^{2}+a_{4}^{\prime }x+a_{6}^{\prime }$\newline and we see that with this choice:\newline - $v_{1},...,v_{t}$ are all places of good reduction for $E^{\prime \prime }. $\newline - $v$ is a place of multiplicative reduction for $E^{\prime \prime }.$% \newline This completes the proof. \end{proof} \medskip \textbf{Acknowledgements: }First of all, I would like to thank my advisor Jan Nekov\'{a}\v{r}, for suggesting to me this topic, for his guidance along the work and his careful reading of the different versions of this paper. I would also like to thank Vladimir and Tim Dokchitser, the first one for his responses to my questions about their articles, the second one for his advice.
\section*{Introduction} \label{sec:gjsis_intro:introduction} The main task of this paper is to overview a series of our results achieved recently in understanding integrability properties of partial differential equations (PDEs) arising in mathematical physics and geometry~\cite{KerstenKrasilshchikVerbovetsky:HOpC,KerstenKrasilshchikVerbovetsky:InCSSREvDEq,KerstenKrasilshchikVerbovetsky:MAmEqHSSRH,KerstenKrasilshchikVerbovetsky:NHSSRHNApApNSKEq,KerstenKrasilshchikVerbovetsky:GSDBTEq,GolovkoKrasilshchikVerbovetsky:InCHEqIn,GolovkoKrasilshchikVerbovetsky:VPNSPDEq,IgoninKerstenKrasilshchikVerbovetskyVitolo:VBGPDE,KerstenKrasilshchikVerbovetskyVitolo:HSGP,KerstenKrasilshchikVerbovetskyVitolo:InKD}. These results are essentially based on the geometrical approach to PDEs developed since the 1970s by A.~Vinogradov and his school (see~\cite{KrasilshchikLychaginVinogradov:GJSNPDEq,KrasilshchikVinogradov:SCLDEqMP,KrasilshchikVerbovetsky:HMEqMP,Vinogradov:CAnPDEqSC,KrasilshchikKersten:SROpCSDE} and references therein). The approach treats a PDE as an (infinite-dimensional) submanifold in the space~$J^\infty(\pi)$ of infinite jets for a bundle~$\pi\colon E\to M$ whose sections play the r\^{o}le of unknown functions (fields). This attitude allowed to apply to PDEs powerful techniques of differential geometry and homological algebra. The latter, in particular, made it possible to give an invariant and efficient formulation of higher-order Lagrangian formalism with constraints and calculus of variations (see~\cite{Vinogradov:SSLFCLLTNT,Vinogradov:AlGFLFT,Vinogradov:CAnPDEqSC,Vinogradov:SSAsNDEqAlGFLFTC}), see also~\cite{Takens:GVInPCV,Anderson:InVB,Tsujishita:VBAsDEq,Tsujishita:HMCInSDEq,Tulczyjew:LC,Kupershmidt:GJBSLHF,Manin:AlAsNDEq} and, as it became clear later, was a bridge to BRST cohomology in gauge theories, anti-field formalism and related topics,~\cite{Henneaux:CInGFCAp}; see also~\cite{BarnichBrandtHenneaux:LBCAnFIGT,HenneauxTeitelboim:QGS,GomisParisSamuel:AnAnGTQ}. Geometrical treatment of differential equations has a long history and originates in the works by Sophus Lie~\cite{Lie:VDT,Lie:GB,Lie:GAb}, as well as in research by A.V.~B\"{a}cklund~\cite{Baecklund:ZTF}, G.~Monge~\cite{Monge:ApG}, G.~Darboux~\cite{Darboux:LTGSApGCIn}, L.~Bianchi~\cite{Bianchi:LGD} and, later, by \'{E}lie Cartan~\cite{Cartan:SDExApG}. Note incidentally that Cartan's theory of involutivity for external differential systems was an inspiration for another cohomological theory associated to PDEs and developed in papers by D.~Spencer and his school,~\cite{Spencer:OvSLPDEq,Goldschmidt:ExTAnLPDEq}. Spencer's work (the so-called formal theory) closely relates to earlier and unfairly forgotten results by M.~Janet~\cite{Janet:LSDP} and Ch.~Riquier~\cite{Riquier:LSDP}; see also~\cite{Pommaret:SPDEqLP} as well as~\cite{Rashevskii:GTPDEq,Seiler:FTDEqApCAl,BryantChernGardnerGoldschmidtGriffiths:ExDS,KruglikovLychagin:GDEq}. A milestone in the geometry of differential equations was introduction by Charles Ehresmann the notion of jet bundles~\cite{Ehresmann:InTSInPL1,Ehresmann:InTSInPL2} that became the most adequate language for Lie's theory, but a real revival of the latter came with the works by L.V.~Ovsyannikov (see his book~\cite{Ovsiannikov:GAnDEq} on group analysis of PDEs; see also~\cite{Ibragimov:TGApMP,Stephani:DEqTSUsS,BlumanKumei:SDEq,BlumanCheviakovAnco:ApSMPDEq}). A new impulse for the reappraisal of the Sophus Lie heritage was given by the discovery of integrability phenomena in nonlinear systems~\cite{GardnerGreeneKruskalMiura:MSKVEq,Miura:KVEqGIRExNT,MiuraGardnerKruskal:KVEqGIIExCLCM,SuGardner:KVEqGIIIDKVEqBEq,Gardner:KVEqGIVKVEqHS,KruskalMiuraGardnerZabusky:KVEqGVUnNPCL,GardnerGreeneKruskalMiura:KEqGVMExS,BulloughCaudrey:S,NovikovManakovPitaevskiiZakharov:TS} in the fall of the 1960s\footnote{Though discussions on ``what is integrability'' continued~\cite{Zakharov:WIsIn} and are still held now (see, e.g. quite recent papers~\cite{FerapontovKhusnutdinovaKlein:LDInQSHD} or~\cite{OdesskiiSokolov:InPRGHF}).} and Hamiltonian interpretation of this integrability~\cite{ZakharovFaddeev:KdVEqisFInHS,Gardner:KVEqGIVKVEqHS}. In particular, it became clear that integrable equations possess infinite series of higher, or generalised, symmetries (see~\cite{Olver:ApLGDEq,KrasilshchikLychaginVinogradov:GJSNPDEq}), and classification of evolution equations with respect to this property allowed to discover new, at that time, integrable equations,~\cite{MikhailovShabatYamilov:SApCNEqCLInS,SvinolupovSokolovYamilov:BTInEvEq,Mikhailov:In}. Later the notion of a higher symmetry was generalised further to that of a nonlocal one~\cite{VinogradovKrasilshchik:MCHSNEvEqNS} and the search for a geometrical background of nonlocality led to the concept of a differential covering~\cite{KrasilshchikVinogradov:NTGDEqSCLBT}. The latter proved to play an important r\^{o}le in the geometry of PDEs and we discuss it in our review. It also became clear that the majority of integrable evolutionary systems possess a bi-Hamiltonian structure~\cite{Magri:SInHP,Magri:SMInHEq}, i.e., can be represented as Hamiltonian flows on the space of infinite jets in at least two different ways and the corresponding Hamiltonian structures are compatible. The bi-hamiltonian property, by Magri's scheme~\cite{Magri:SMInHEq}, leads to the existence of infinite series of commuting symmetries and conservation laws. In addition, it gives rise to a recursion operator for higher symmetries that is an efficient tool for practical construction of symmetry hierarchies. Nevertheless, recursion operators exist for equations possessing no Hamiltonian structure at all (e.g., for the Burgers equation). A self-contained cohomological approach to recursion operators based on Nijenhuis brackets and related to the theory of deformations for PDE structures is exposed in~\cite{KrasilshchikKersten:SROpCSDE}. The literature on the Hamiltonian theory of PDEs is vast and we confine ourselves here to the key references~\cite{FaddeevTakhtadzhyan:HMTS,Dickey:SEqHS,Dorfman:DSInNEvEq,DubrovinKricheverNovikov:InSI}, but one feature is common to all research: theories and techniques are applicable to evolution equations only. Then a natural question arises: what to do if the equation at hand is not represented in the evolutionary form? We believe that (at least, a partial) answer to this question can be found in this paper. Of course, one of possible solutions is to transform the equation to the evolutionary form. But: \begin{itemize} \item Not all equations can be rendered to this form\footnote{For example, gauge invariant equations, such the Yang-Mills, Maxwell, Einstein equation, etc., can not be presented in the evolutionary form.}. \item How to check independence of Hamiltonian (and other) structures on a particular representation of our equation? In other words, if we found a Hamiltonian operator in one representation what guarantees that it survives when the representation is changed? \item Even if the answer to the previous question is positive, how to transform the results when passing to the initial form of the equation? \end{itemize} In what follows, we treat any concrete equation ``as is'' and try to uncover those objects and constructions that are naturally associated to this equation. In particular, we do not assume existence of any additional structures that enrich the equation. Such structures are by all means extremely interesting and lead to very nontrivial classes of equations (e.g., equations of hydrodynamical type~\cite{DubrovinNovikov:PBHT}, Monge-Amp\'{e}re equations~\cite{KushnerLychaginRubtsov:CGNLDEq} or equations associated to Lie groups~\cite{DrinfeldSokolov:LAlEqKVT}), but here we look for internal properties of an arbitrary PDE. As it was said in the very beginning, an equation (or, to be more precise, its infinite prolongation, i.e., equation itself together with all its differential consequences) is a submanifold in a jet space~$J^\infty(\pi)$. To escape technical difficulties, we consider the simplest case, when~$\pi\colon E\to M$ is a locally trivial vector bundle, though all the results remain valid in a more complicated situation (e.g., for jets of submanifolds). The reader who is interested in local results only may keep in mind the trivial bundle~$\mathbb{R}^m\times\mathbb{R}^n\to\mathbb{R}^n$ instead of~$\pi$. Understood in such a way, any equation~$\mathcal{E}\subsetJ^\infty(\pi)$ is naturally endowed with a $(\dim M)$-dimensional integrable distribution~$\mathcal{C}$ (the Cartan distribution) which consists, informally speaking, of planes tangent to formal solutions of~$\mathcal{E}$. This is the main and essentially the only geometric structure that we use. In the research of the PDE differential geometry we use somewhat informal but quite productive guidelines which were originally introduced in~\cite{Vinogradov:CNDEq} (see also~\cite{Vinogradov:InInGJS,Vinogradov:GNDEq}) and may be formulated as \begin{description} \item[The structural principle] Any construction and concept must take into account the Cartan distribution on~$\mathcal{E}$. \item[The correspondence principle] ``Physical dimension'' of~$\mathcal{E}$ is~$n=\dim\mathcal{C}=\dim M$ and differential geometry of~$\mathcal{E}$ reduces to the finite-dimensional one when passing to the ``classical limit~$n\to 0$''. \item[The invariance principle] All constructions must be independent of the embedding~$\mathcal{E}\toJ^\infty(\pi)$ and defined by the equation~$\mathcal{E}$ itself. \end{description} Below we accompany our exposition by toy dictionaries that illustrate the correspondence between two languages, those of the geometry of PDEs and classical differential geometry. The paper consists of three sections. In Section~\ref{sec:gjsis_jets:jet-spaces} we describe the geometry of the ``empty equation'', i.e., of the jet space~$J^\infty(\pi)$. In particular, we define the tangent and cotangent bundles to~$J^\infty(\pi)$, introduce variational differential forms and multivectors and define the variational Schouten bracket. We discuss geometry of Hamiltonian flows on the space of infinite jets (i.e., Hamiltonian evolutionary equations) and Lagrangian formalism without constraints. Section~\ref{sec:gjsis_eqs:diff-equat} deals with the same matters, but in the context of a differential equation~$\mathcal{E}\subsetJ^\infty(\pi)$. Although the exposition in this part is quite general, the result on the Hamiltonian theory (the definition of the cotangent bundle, in particular) are valid for the so-called $2$-line equations only (we call such equations normal in Section~\ref{sec:gjsis_eqs:diff-equat}). This notion is related to the cohomological length of the compatibility complex for the linearization operator of~$\mathcal{E}$ and manifests itself, for example, in the number of nontrivial lines in Vinogradov's $\mathcal{C}$-spectral sequence,~\cite{Vinogradov:SSLFCLLTNT} (see also~\cite{Vinogradov:SSAsNDEqAlGFLFTC,Vinogradov:CAnPDEqSC,BryantGriffiths:CCDSGT} and~\cite{KrasilshchikVerbovetsky:HMEqMP,Verbovetsky:NHC}). From this point of view, jet spaces are $1$-line equations and this is the reason why they have to be treated separately. Finally, in Section~\ref{sec:gjsis_nonloc:nonlocal-theory} we briefly overview the theory of differential coverings for PDEs and some of its applications: nonlocal symmetries and shadows, B\"{a}cklund transformations, etc. Our exposition of general facts related to the geometry of jet spaces and infinitely prolonged equations, including the nonlocal theory, is essentially based on the books~\cite{KrasilshchikLychaginVinogradov:GJSNPDEq,KrasilshchikVinogradov:SCLDEqMP,KrasilshchikVerbovetsky:HMEqMP}. Lagrangian formalism, both in the free case (on jets) and with constraints (on equations), is exposed using the material of~\cite{Vinogradov:CAnPDEqSC} (see also~\cite{KrasilshchikVinogradov:SCLDEqMP,KrasilshchikVerbovetsky:HMEqMP}). The geometrical approach to Hamiltonian formalism (including the theory of the Schouten bracket) is based on~\cite{KerstenKrasilshchikVerbovetsky:HOpC,KerstenKrasilshchikVerbovetskyVitolo:HSGP}. A practical implementation of the general theory, in the majority of cases, needs the use of an appropriate computer algebra software. To avoid technical details that obscure the essentials, we chose for a ``tutorial example'' the well known Korteweg-de Vries equation for which all computations are transparent and can be done by hand. We did our best to illustrate the theory by a reasonable number of less trivial examples and really do hope that the result will be interesting to the readers. \section{Jet spaces} \label{sec:gjsis_jets:jet-spaces} Jet spaces constitute a natural geometric environment for differential equations and for equations of mathematical physics, in particular. But these spaces are themselves an interesting geometric object that contains information on Lagrangian and Hamiltonian formalisms without constraints. Thus, we begin our exposition with a description of these spaces and structures related to them. \subsection{Definition of jet spaces} \label{sec:gjsis_jets:defin-jet-spac} Let $\pi\colon E\to M$ be a locally trivial smooth vector bundle\footnote{For a definition of jets in a more general setting see, e.g.,~\cite{KrasilshchikLychaginVinogradov:GJSNPDEq,Vinogradov:CAnPDEqSC,Saunders:GJB,Saunders:JMNB}.} over a smooth manifold~$M$, $\dim M=n$, $\dim E=m+n$. In what follows, $M$ will be the manifold of independent variables while sections of~$\pi$ will play the r\^{o}le of unknown functions (fields). The set of all sections~$s\colon M\to E$ will be denoted by~$\Gamma(\pi)$ and it forms a module over the algebra~$C^\infty(M)$. Two sections~$s$, $s'\in\Gamma(\pi)$ are said to be \emph{$k$-equivalent} at a point~$x\in M$ if their graphs are tangent to each other with order~$k$ at the point~$s(x)=s'(x)\in E$. The equivalence class of~$s$ with respect to this relation is denoted by~$[s]_x^k$ and is called the \emph{$k$-jet} of~$s$ at~$x$. The set \begin{equation*} J^k(\pi)=\sd{[s]_x^k}{x\in M,s\in\Gamma(\pi)} \end{equation*} is endowed with a natural structure of a smooth manifold; the latter is called the \emph{manifold of $k$-jets} of sections of~$\pi$. Moreover, the maps \begin{equation}\label{eq:gjsis_jets:1} \pi_k\colonJ^k(\pi)\to M,\quad [s]_x^k\mapsto x, \end{equation} and \begin{equation}\label{eq:gjsis_jets:2} \pi_{k,l}\colonJ^k(\pi)\toJ^l(\pi),\quad [s]_x^k\mapsto[s]_x^l,\qquad k\ge l, \end{equation} are smooth fibre bundles, $\pi_k$ being vector bundles. For any section~$s\in\Gamma(\pi)$ the map \begin{equation}\label{eq:gjsis_jets:3} j_k(s)\colon M\toJ^k(\pi),\quad x\mapsto[s]_x^k, \end{equation} is a smooth section of~$\pi_k$ that is called the \emph{$k$-jet} of~$s$. Here we are mostly interested in the case~$k=\infty$, i.e., in the space~$J^\infty(\pi)$. It can be understood as the inverse limit of the chain \begin{equation}\label{eq:gjsis_jets:4} \qquad\cdots\xrightarrow{}J^{k+1}(\pi)\xrightarrow{\pi_{k+1,k}}J^k(\pi)\xrightarrow{}\cdots\xrightarrow{} J^1(\pi)\xrightarrow{\pi_{1,0}}J^0(\pi)=E\xrightarrow{\pi}M. \end{equation} Due to projections~\eqref{eq:gjsis_jets:4} there exist monomorphisms of function algebras \begin{equation}\label{eq:gjsis_jets:5} C^\infty(M)\subset\mathcal{F}_0(\pi)\subset\dots\subset\mathcal{F}_k(\pi)\subset\mathcal{F}_{k+1}(\pi) \subset\dots, \end{equation} where~$\mathcal{F}_k(\pi)=C^\infty(J^k(\pi))$, and we define the \emph{algebra of smooth functions} on~$J^\infty(\pi)$ as the filtered algebra~$\mathcal{F}(\pi)=\cup_k\mathcal{F}_k(\pi)$. Elements of~$\mathcal{F}(\pi)$ are identified with nonlinear scalar differential operators acting on sections of~$\pi$ by the following rule: \begin{equation}\label{eq:gjsis_jets:6} \Delta_f(s)=j_\infty(s)^*(f),\qquad s\in\Gamma(\pi),f\in\mathcal{F}(\pi). \end{equation} More general, let~$\pi'\colon E'\to M$ be another vector bundle and~$\pi^*(\pi')$ be its pull-back to~$J^\infty(\pi)$. Introduce the notation~$\mathcal{F}(\pi,\pi')=\Gamma(\pi^*(\pi'))$. Then any section~$f\in\mathcal{F}(\pi,\pi')$ is identified, by a formula similar to~\eqref{eq:gjsis_jets:6}, with a nonlinear differential operator that acts from~$\Gamma(\pi)$ to~$\Gamma(\pi')$. \subsection{Vector fields and differential forms} \label{sec:gjsis_jets:vect-fields-diff} A \emph{vector field} on~$J^\infty(\pi)$ is a derivation of the function algebra~$\mathcal{F}(\pi)$, i.e., an $\mathbb{R}$-linear map~$X\colon\mathcal{F}(\pi)\to\mathcal{F}(\pi)$ such that \begin{equation*} X(fg)=fX(g)+gX(f) \end{equation*} for all $f$, $g\in\mathcal{F}(\pi)$. The set of all vector fields is denoted by~$\mathcal{X}(\pi)$ and it is a Lie algebra with respect to the commutator (the \emph{Lie bracket}). The definition of a \emph{differential form} of degree~$r$ on~$J^\infty(\pi)$ is similar to that of smooth functions. Using projections~\eqref{eq:gjsis_jets:4} we consider the embeddings~$\Lambda^r(J^k(\pi))\subset\Lambda^r(J^{k+1}(\pi))$ and set~$\Lambda^r(\pi)=\cup_k\Lambda^r(J^k(\pi))$. We shall also consider the Grassmann algebra of all forms~$\Lambda^*(\pi)=\oplus_{r\ge 0}\Lambda^r(\pi)$ with respect to the wedge product. \begin{coordinates} \label{sec:gjsis_jets:vect-fields-diff-1} Let~$\mathcal{U}\subset M$ be a coordinate neighbourhood such that the bundle~$\pi$ becomes trivial over~$\mathcal{U}$. Choose local coordinates~$x^1,\dots,x^n$ in~$\mathcal{U}$ and~$u^1,\dots,u^m$ along the fibres of~$\pi$ over~$\mathcal{U}$. Then the \emph{adapted} coordinates in~$\pi^{-1}(\mathcal{U})\subsetJ^\infty(\pi)$ naturally arise. These coordinates are denoted by~$u_I^j$, $I$ being a multi-index, and are defined by \begin{equation*} j_\infty(s)^*(u_I^j)=\frac{\partial^{\abs{I}}s^j}{\partial x^I}, \end{equation*} where~$s=(s^1,\dots,s^m)$ is a local section of~$\pi$ over~$\mathcal{U}$. In other words, the coordinate functions~$u_I^j$ correspond to partial derivatives of local sections. In these coordinates, smooth function on~$J^\infty(M)$ are of the form \begin{equation*} f=f(x^i,u_I^j), \end{equation*} where the number of arguments is \emph{finite}. Vector fields are represented as \emph{infinite} sums \begin{equation*} X=\sum_ia_i\frac{\partial}{\partial x^i}+ \sum_{I,j}a_I^j\frac{\partial}{\partial u_I^j},\qquad a_i,a_I^j\in\mathcal{F}(\pi), \end{equation*} while differential forms of degree~$r$ are \emph{finite} sums \begin{equation*} \omega=\sum b_{i_1,\dots,i_c,j_{c+1},\dots,j_r}^{I_{c+1},\dots,I_r} \mathinner{\!}\mathrm{d} x^{i_1}\wedge\dots\wedge\mathinner{\!}\mathrm{d} x^{i_c}\wedge \mathinner{\!}\mathrm{d} u_{I_{c+1}}^{j_{c+1}}\wedge\dots\wedge\mathinner{\!}\mathrm{d} u_{I_r}^{j_r}. \end{equation*} \end{coordinates} \subsection{Main structure: the Cartan distribution} \label{sec:gjsis_jets:main-struct-cart} Let~$\theta\inJ^\infty(\pi)$. Then the graphs of all sections~$j_\infty(s)$, $s\in\Gamma(\pi)$, passing through the point~$\theta$ have a common $n$-dimensional tangent plane~$\mathcal{C}_\theta$ (the \emph{Cartan plane}). The correspondence~$\mathcal{C}\colon\theta\mapsto\mathcal{C}_\theta$ is an integrable\footnote{\label{fn:1}Integrability is understood formally here and means that if~$X$ and~$Y$ are two vector fields lying in~$\mathcal{C}$ then their bracket~$[X,Y]$ lies in~$\mathcal{C}$ as well. Since~$J^\infty(\pi)$ is infinite-dimensional, this does not mean that the Frobenius theorem holds for~$\mathcal{C}$: for any point~$\theta\inJ^\infty(\pi)$ there exist infinitely many maximal integral manifolds that contain~$\theta$. On the other hand, if~$\mathcal{C}$ is restricted to an equation (see below Section~\ref{sec:gjsis_eqs:diff-equat}), there may exist no maximal integral manifold at all.} $n$-dimensional distribution on~$J^\infty(\pi)$ that is called the \emph{Cartan distribution}. This distribution is the basic geometric structure on the manifold~$J^\infty(\pi)$. In particular, the following result is valid: \begin{proposition} \label{prop:gjsis_jets:1} A submanifold in~$J^\infty(M)$ is a maximal integral manifold of~$\mathcal{C}$ if and only if it is the graph of~$j_\infty(s)$, where~$s$ is a local section of~$\pi$. \end{proposition} Moreover, since the planes~$\mathcal{C}_\theta$ project to~$M$ non-degenerately, any vector field~$X$ on~$M$ can be uniquely lifted up to a field~$\mathcal{C} X$ on~$J^\infty(\pi)$. In such a way, one obtains a connection in the bundle~$\pi_\infty$ called the \emph{Cartan connection}. This connection is flat, i.e., \begin{equation}\label{eq:gjsis_jets:7} \mathcal{C}[X,Y]=[\mathcal{C} X,\mathcal{C} Y] \end{equation} for all vector fields~$X$, $Y$ on~$M$. Due to \eqref{eq:gjsis_jets:7}, the space~$\mathcal{C}\mathcal{X}(\pi)$ of all vector fields lying in the Cartan distribution is a Lie subalgebra in~$\mathcal{X}(\pi)$. Vector fields belonging to~$\mathcal{X}(\pi)$ will be called \emph{Cartan fields}. Any vector field~$Z\in\mathcal{X}(\pi)$ can be uniquely decomposed to its \emph{vertical} and \emph{horizontal} components, \begin{equation}\label{eq:gjsis_jets:8} Z=Z^v+Z^h, \end{equation} where~$Z^v$ is the projection of~$X$ to the fibre of the bundle~$\pi_\infty$ along Cartan planes, while~$Z^h$ lies in the Cartan distribution. Thus, one has \begin{equation}\label{eq:gjsis_jets:9} \mathcal{X}(\pi)=\mathcal{X}^v(\pi)\oplus\mathcal{C}\mathcal{X}(\pi), \end{equation} where~$\mathcal{X}^v(\pi)$ is the Lie algebra of vertical vector fields. Dually to~\eqref{eq:gjsis_jets:9}, the module of differential forms~$\Lambda^1(\pi)$ splits into the direct sum \begin{equation}\label{eq:gjsis_jets:10} \Lambda^1(\pi)=\La_{\scriptscriptstyle{\mathcal{C}}}^1(\pi)\oplus\La_h^1(\pi), \end{equation} where~$\La_{\scriptscriptstyle{\mathcal{C}}}^1(\pi)$ consists of $1$-forms that annihilate the Cartan distribution (they will be called \emph{Cartan forms}, or \emph{higher contact forms}), while elements of~$\La_h^1(\pi)$ are \emph{horizontal forms}. \begin{coordinates} \label{sec:gjsis_jets:main-struct-cart-1} Choose an adapted coordinate system~$(x^i,u_I^j)$ in~$J^\infty(\pi)$. Then one has \begin{equation}\label{eq:gjsis_jets:11} \mathcal{C}\colon\frac{\partial}{\partial x^i}\mapsto D_i= \frac{\partial}{\partial x^i}+ \sum_{I,j}u_{Ii}^j\frac{\partial}{\partial u_I^j}. \end{equation} The fields~$D_i$ are called \emph{total derivatives} and they span the Cartan distribution. For a basis in the module~$\La_{\scriptscriptstyle{\mathcal{C}}}^1(\pi)$ one can choose the forms \begin{equation}\label{eq:gjsis_jets:12} \omega_I^j=\mathinner{\!}\mathrm{d} u_I^j-\sum_i u_{Ii}^j\mathinner{\!}\mathrm{d} x^i, \end{equation} while horizontal forms are \begin{equation}\label{eq:gjsis_jets:13} \omega=\sum_i a_i\mathinner{\!}\mathrm{d} x^i,\qquad a_i\in\mathcal{F}(\pi). \end{equation} \end{coordinates} \begin{remark} \label{rem:gjsis_jets:6} It should be noted that all results and constructions below are valid not for the entire jet space only, but for an arbitrary open domain in~$J^\infty(\pi)$. Everywhere below, when speaking about~$J^\infty(\pi)$, we actually mean an open domain. \end{remark} \subsection{Evolutionary vector fields and linearizations} \label{sec:gjsis_jets:evol-vect-fields} We shall now describe infinitesimal symmetries of the Cartan distribution on~$J^\infty(\pi)$. A vector field~$X\in\mathcal{X}(\pi)$ is a \emph{symmetry} if~$[X,Z]\in\mathcal{C}\mathcal{X}(\pi)$ as soon as~$Z\in\mathcal{C}\mathcal{X}(\pi)$. The space~$\mathcal{C}\mathcal{X}(\pi)$ is an ideal in the Lie algebra~$\X_{\scriptscriptstyle{\mathcal{C}}}(\pi)$ of symmetries. Due to integrability of the Cartan distribution, any~$Z\in\mathcal{C}\mathcal{X}(\pi)$ is a symmetry, and we call such symmetries \emph{trivial}. Thus, we introduce the Lie algebra of \emph{nontrivial symmetries} as \begin{equation*} \sym\pi=\X_{\scriptscriptstyle{\mathcal{C}}}(\pi)/\mathcal{C}\mathcal{X}(\pi). \end{equation*} By \eqref{eq:gjsis_jets:9},~$\sym\pi$ is identified with the vertical part of~$\X_{\scriptscriptstyle{\mathcal{C}}}(\pi)$. Take a vector field~$X\in\sym\pi$ and restrict it to the subalgebra~$\mathcal{F}_0(\pi)\subset\mathcal{F}(\pi)$. Then this restriction can be identified with an element~$\phi_X\in\mathcal{F}(\pi,\pi)$. For shortness, we shall use the notation~$\mathcal{F}(\pi,\pi)=\kappa(\pi)$. \begin{theorem} \label{thm:gjsis_jets:1} The correspondence~$X\mapsto\varphi_X$ defines a bijection between~$\sym\pi$ and~$\kappa(\pi)$. \end{theorem} The element~$\phi_X$ is called the \emph{generating section} of a symmetry~$X$, while the symmetry corresponding to a section~$\phi\in\kappa(\pi)$ is called an \emph{evolutionary vector field} and is denoted by~$\bi{E}_\phi$. Theorem~\ref{thm:gjsis_jets:1} allows to introduce an $\mathcal{F}(\pi)$-module structure into the Lie algebra~$\sym\pi$ by setting \begin{equation*} f\cdot\bi{E}_\phi=\bi{E}_{f\phi}. \end{equation*} This multiplication differs from the usual multiplication of a vector field by functions and does not survive when passing from the space of jets to equations (see Section~\ref{sec:gjsis_eqs:diff-equat} below). On the other hand, the same theorem defines a Lie algebra structure in~$\kappa(\pi)$: the \emph{Jacobi bracket}~$\{\phi,\psi\}$ is uniquely given by the equality \begin{equation}\label{eq:gjsis_jets:14} \bi{E}_{\{\phi,\psi\}}=[\bi{E}_\phi,\bi{E}_\psi]. \end{equation} The Jacobi bracket can also be computed using the formula \begin{equation}\label{eq:gjsis_jets:15} \{\phi,\psi\}=\bi{E}_\phi(\psi)-\bi{E}_\psi(\phi). \end{equation} \begin{remark}\label{sec:gjsis_jets:Lie-Back} We pointed out in Footnote~\ref{fn:1} that integrability of distributions on~$J^\infty(\pi)$ differs from the one on finite-dimensional manifolds. The same holds for integrability (i.e., existence of the corresponding one-parameter group of transformations) of vector fields. Generally speaking, a vector field on~$J^\infty(\pi)$ is not integrable in this sense, but there exists an important class of vector fields that are integrable. Namely, if~$X$ is a field on~$J^k(\pi)$, $k<\infty$, that preserves the Cartan distribution then it can be \emph{lifted} in a natural way to~$J^{k+1}(\pi)$, see~\cite{KrasilshchikVinogradov:SCLDEqMP}. The entire collection of such fields determines a vector field on~$J^\infty(\pi)$, and any such a field possesses a one-parameter group of transformations. The infinitesimal version of the Lie-B\"acklund theorem states that all such fields are the lifts of arbitrary fields on~$J^0(\pi)$ (when~$\dim\pi>1$) or of a contact vector field on~$J^1(\pi)$ (when~$\dim\pi=1$). A complete description of integrable vector fields on~$J^\infty(\pi)$ can be found in~\cite{Chetverikov:SInFInPEq,Chetverikov:SInF}. \end{remark} \begin{coordinates} \label{sec:gjsis_jets:evol-vect-fields-1} Let, in an adapted coordinate system, a section~$\phi\in\kappa(\pi)$ be of the form~$\phi=(\phi^1,\dots,\phi^m)$. Then the corresponding evolutionary vector field is \begin{equation}\label{eq:gjsis_jets:16} \bi{E}_\phi=\sum_{I,j}D_I(\phi^j)\frac{\partial}{\partial u_I^j}, \end{equation} where~$D_I=D_{i_1}\circ\dots\circ D_{i_l}$ is the composition of total derivatives corresponding to the multi-index~$I=i_1\dots i_l$. If~$\psi=(\psi^1,\dots,\psi^m)$ is another element of~$\kappa(\pi)$ then the components of the Jacobi bracket are \begin{equation}\label{eq:gjsis_jets:17} \{\phi,\psi\}^j= \sum_{I,\alpha}\left(D_I(\phi^\alpha)\frac{\partial\psi^j}{\partial u_I^\alpha} -D_I(\psi^\alpha)\frac{\partial\phi^j}{\partial u_I^\alpha}\right),\qquad j=1,\dots,m. \end{equation} \end{coordinates} Fix a section~$\psi\in\kappa(\pi)$ and consider the map \begin{equation}\label{eq:gjsis_jets:18} \ell_\psi\colon\kappa(\pi)\to\kappa(\pi),\qquad \ell_\psi(\phi)=\bi{E}_\phi(\psi). \end{equation} This map is called the \emph{linearization} of the element~$\psi$ (recall that~$\psi$ may be identified with a nonlinear differential operator acting from~$\pi$ to~$\pi$). More generally, let~$\pi'\colon E'\to M$ be a vector bundle. Then the action of an evolutionary vector field~$\bi{E}_\phi$ can be extended to \begin{equation*} \bi{E}_\phi\colon\mathcal{F}(\pi,\pi')\to\mathcal{F}(\pi,\pi') \end{equation*} in a well defined way. Consider a section~$\psi\in\mathcal{F}(\pi,\pi')$, i.e., a nonlinear differential operator from~$\Gamma(\pi)$ to~$\Gamma(\pi')$. Its linearization is the map \begin{equation}\label{eq:gjsis_jets:19} \ell_\psi\colon\kappa(\pi)\to\mathcal{F}(\pi,\pi') \end{equation} is defined similar to~\eqref{eq:gjsis_jets:18}. \begin{coordinates}\label{sec:gjsis_jets:evol-vect-fields-2} Let, in adapted coordinates,~$\phi=(\phi^1,\dots,\phi^m)$ and~$\psi=(\psi^1,\dots,\psi^{m'})$. Then the $j$th component of~$\ell_\psi(\phi)$ is \begin{equation*} \sum_{I,\alpha}\frac{\partial\psi^j}{\partial u_I^\alpha}D_I(\phi^\alpha), \end{equation*} i.e., the linearization is a matrix operator of the form \begin{equation}\label{eq:gjsis_jets:20} \ell_\psi= \left\Vert\sum_I\frac{\partial\psi^j}{\partial u_I^\alpha}D_I \right\Vert_{\alpha=1,\dots,m.}^{j=1,\dots,m'} \end{equation} Using linearizations, formula~\eqref{eq:gjsis_jets:15} can be rewritten as \begin{equation}\label{eq:gjsis_jets:21} \{\phi,\psi\}=\ell_\psi(\phi)-\ell_\phi(\psi). \end{equation} \end{coordinates} \subsection{$\mathcal{C}$-differential operators} \label{sec:gjsis_jets:c-diff-oper} From \eqref{eq:gjsis_jets:20} we see that linearizations are differential operators in total derivatives. We call such operators \emph{$\mathcal{C}$-differential operators}. More precisely, let~$\xi$ and~$\xi'$ be two vector bundles over~$J^\infty(\pi)$ and~$P$, $P'$ be the $\mathcal{F}(\pi)$-modules of their sections. An $\mathbb{R}$-linear map~$\Delta\colon P\to P'$ is a $\mathcal{C}$-differential operator of order~$k$ if for any point~$\theta\inJ^\infty(\pi)$ and a section~$p\in P$ the value of~$\Delta(p)$ at~$\theta$ is completely determined by the values of~$D_I(p)$, $\abs{I}\le k$, at this point. The space of all such operators is denoted by~$\CDiff_k(P,P')$ and we also set~$\CDiff(P,P')=\cup_k\CDiff_k(P,P')$. A closely related notion to that of a $\mathcal{C}$-differential operator is \emph{horizontal jets}. Let~$P$ be as above. We say that two sections~$p$, $p'\in P$ are \emph{horizontally $k$-equivalent} (the case~$k=\infty$ is included) at a point~$\theta\inJ^\infty(\pi)$ if~$D_I(p)=D_I(p')$ at~$\theta$ for all~$I$ such that~$\abs{I}\le k$. Denote the equivalence class by~$\{p\}_\theta^k$. The set \begin{equation*} J_h^k(P)=\sd{\{p\}_\theta^k}{\theta\inJ^\infty(\pi),p\in P} \end{equation*} forms a smooth manifold which is fibred over~$J^\infty(\pi)$, \begin{equation*} \xi_k\colonJ_h^k(P)\toJ^\infty(\pi)\quad\{\rho\}_\theta^k\mapsto\theta. \end{equation*} The section~$j_k^h(p)\colonJ^\infty(\pi)\toJ_h^k(P)$, $\theta\mapsto\{p\}_\theta^k$, is called the \emph{horizontal jet} of~$p\in P$. \begin{proposition} \label{prop:gjsis_jets:2} For any $\mathcal{C}$-differential operator~$\Delta\in\CDiff_k(P,P')$ there exists a unique morphism~$\Phi_\Delta$ of vector bundles~$\xi_k$ and~$\xi'$ such that~$\Delta(p)=\Phi_\Delta(j_k^h(p))$ for any~$p\in P$. \end{proposition} Two natural identifications will be useful below. \begin{proposition} \label{prop:gjsis_jets:3} For any vector bundle~$\pi$ one has: \begin{enumerate} \item The module~$\La_{\scriptscriptstyle{\mathcal{C}}}^1(\pi)$ is isomorphic to~$\CDiff(\kappa(\pi),\mathcal{F}(\pi))$. \item The module~$J_h^\infty(\kappa(\pi))$ is isomorphic to~$\mathcal{X}^v(\pi)$. The vector fields corresponding to sections of the form~$j_\infty^h(p)$ are evolutionary fields. \end{enumerate} \end{proposition} \begin{coordinates} \label{sec:gjsis_jets:c-diff-oper-1} Choose an adapted coordinate system in the manifold~$J^\infty(\pi)$ and let~$r$, $r'$ be dimensions of the bundles~$\xi$, $\xi'$, respectively. Then any operator~$\Delta\in\CDiff(P,P')$ is of the form \begin{equation}\label{eq:gjsis_jets:22} \Delta= \left\Vert\sum_Ia_{\alpha\beta}^ID_I\right\Vert_{\alpha=1,\dots,r',}^{\beta=1,\dots,r} \qquad a_{\alpha\beta}^I\in\mathcal{F}(\pi). \end{equation} If~$v^1,\dots,v^r$ are fibre-wise coordinates in the bundle~$\xi$ then the adapted coordinates~$v_K^l$ in~$J_h^\infty(\xi)$, $K$ being a multi-index, $l=1,\dots,r$, are determined by the equalities \begin{equation}\label{eq:gjsis_jets:23} j_\infty^h(s)^*(v_K^l)=D_K(s^l), \end{equation} where~$s=(s^1,\dots,s^r)$ is a local section of the bundle~$\xi$. \end{coordinates} \begin{remark} \label{rem:gjsis_jets:1} The space of horizontal jets~$J_h^\infty(P)$ is also endowed with an integrable distribution similar to the Cartan one: if~$\theta\inJ_h^\infty(P)$ then the corresponding plane~$\mathcal{C}_\theta$ is tangent to the graphs of horizontal jets passing through this point. The differential of the map~$\xi_\infty\colonJ_h^\infty(P)\toJ^\infty(\pi)$ isomorphically projects~$\mathcal{C}_\theta$ to~$\mathcal{C}_{\xi_\infty(\theta)}$. Moreover, if~$P$ is of the form~$P=\Gamma(\pi_\infty^*(\xi))$, where~$\xi$ is a vector bundle over~$M$ and~$\pi_\infty^*(\xi)$ is its pull-back, then one has a diffeomorphism \begin{equation*} J_h^\infty(P)=J^\infty(\pi\times_M\xi), \end{equation*} where~$\pi\times_M\xi$ is the Whitney product, and this isomorphism takes the Cartan distribution on~$J_h^\infty(P)$ to the one on~$J^\infty(\pi\times_M\xi)$. \end{remark} \begin{remark} \label{rem:gjsis_jets:2} In the case when the modules~$P=\Gamma(\pi_\infty^*(\xi))$ and~$P'=\Gamma(\pi_\infty^*(\xi'))$ are of the form considered in Remark~\ref{rem:gjsis_jets:1} $\mathcal{C}$-differential operators from~$P$ to~$P'$ may be understood as non-linear differential operators that take sections of~$\pi$ to linear differential operators acting from~$\xi$ to~$\xi'$. \end{remark} \subsection{The variational complex and Lagrangian formalism} \label{sec:gjsis_jets:vari-compl-lagr} Consider now decomposition~\eqref{eq:gjsis_jets:10} and note that it implies a more general splitting \begin{equation}\label{eq:gjsis_jets:24} \Lambda^k(\pi)=\bigoplus_{p+q=k}\La_{\scriptscriptstyle{\mathcal{C}}}^p(\pi)\otimes\La_h^q(\pi), \end{equation} where \begin{equation*} \La_{\scriptscriptstyle{\mathcal{C}}}^p(\pi)= \underbrace{\La_{\scriptscriptstyle{\mathcal{C}}}^1(\pi)\wedge\dots\wedge\La_{\scriptscriptstyle{\mathcal{C}}}^1(\pi)}_{p\text{ times}},\qquad \La_h^q(\pi)= \underbrace{\La_h^1(\pi)\wedge\dots\wedge\La_h^1(\pi)}_{q\text{ times}}. \end{equation*} Introduce the notation~$\La_{\scriptscriptstyle{\mathcal{C}}}^p(\pi)\otimes\La_h^q(\pi)=E_0^{p,q}(\pi)$. By Proposition~\ref{prop:gjsis_jets:3}~(1), this space is identified with the module~$\CDiff^{\mathrm{sk}}_p(\kappa(\pi),\La_h^q(\pi))$ of $p$-linear skew-symmetric $\mathcal{C}$-differential operators acting from~$\kappa(\pi)$ to~$\La_h^q(\pi)$. The de~Rham differential~$\mathinner{\!}\mathrm{d}\colon\Lambda^k(\pi)\to\Lambda^{k+1}(\pi)$, by~\eqref{eq:gjsis_jets:24}, splits into two parts~$\mathinner{\!}\mathrm{d}=\rmd_{\scriptscriptstyle{\mathcal{C}}}+\rmd_h$, where \begin{equation}\label{eq:gjsis_jets:25} \rmd_{\scriptscriptstyle{\mathcal{C}}}=\rmd_{\scriptscriptstyle{\mathcal{C}}}^{p,q}\colon E_0^{p,q}(\pi)\to E_0^{p+1,q}(\pi),\qquad \rmd_h=\rmd_h^{p,q}\colon E_0^{p,q}(\pi)\to E_0^{p,q+1}(\pi) \end{equation} are the \emph{vertical} (or \emph{Cartan}) and \emph{horizontal} differentials, respectively These differentials anti-commute, i.e., \begin{equation}\label{eq:gjsis_jets:26} \rmd_{\scriptscriptstyle{\mathcal{C}}}\circ\rmd_h+\rmd_h\circ\rmd_{\scriptscriptstyle{\mathcal{C}}}=0. \end{equation} \begin{coordinates} For a function on~$J^\infty(\pi)$ (i.e., a $0$-form) the action of the Cartan differential is given by \begin{equation}\label{eq:gjsis_jets:27} \rmd_{\scriptscriptstyle{\mathcal{C}}} f=\sum_{I,j}\frac{\partial f}{\partial u_I^j}\rmd_{\scriptscriptstyle{\mathcal{C}}} u_I^j, \end{equation} where~$\rmd_{\scriptscriptstyle{\mathcal{C}}} u_I^j=\omega_I^j$ are the Cartan forms presented in \eqref{eq:gjsis_jets:12}, while the horizontal differential acts as follows \begin{equation}\label{eq:gjsis_jets:28} \rmd_h f=\sum_i D_i(f)\mathinner{\!}\mathrm{d} x^i. \end{equation} To compute the action on arbitrary forms it suffices to use~\eqref{eq:gjsis_jets:27} and~\eqref{eq:gjsis_jets:28} and the fact that~$\rmd_{\scriptscriptstyle{\mathcal{C}}}$ and~$\rmd_h$ differentiate the wedge product and anti-commute with the de~Rham differential. \end{coordinates} Let~$E_1^{p,q}(\pi)$ be the cohomology of~$\rmd_h$ at the term~$E_0^{p,q}(\pi)$, i.e., \begin{equation}\label{eq:gjsis_jets:29} E_1^{p,q}(\pi)=\ker\rmd_h^{p,q}/\im\rmd_h^{p,q-1}. \end{equation} Due to~\eqref{eq:gjsis_jets:26}, the vertical differential~$\rmd_{\scriptscriptstyle{\mathcal{C}}}$ induces the differentials~$E_1^{p,q}(\pi)\to E_1^{p+1,q}(\pi)$ which will be denoted by~$\delta^{p,q}$. The groups~$E_1^{p,q}(\pi)$, together with the differentials~$\delta^{p,q}$, play one of the most important r\^{o}les in the geometry of jets providing the background for Lagrangian formalism without constraints. To describe them, we shall need new notions. Let~$\xi$ be a vector bundle over~$J^\infty(\pi)$ and~$P=\Gamma(\xi)$. Introduce the \emph{adjoint} module $\hat{P}=\Hom_{\mathcal{F}(\pi)}(P,\La_h^n(\pi))$. Consider another module of sections~$Q$ and a $\mathcal{C}$-differential operator~$\Delta\colon P\to Q$. Then the \emph{adjoint} operator~$\Delta^*\colon\hat{Q}\to\hat{P}$ is defined and it enjoys the \emph{Green formula} \begin{equation}\label{eq:gjsis_jets:30} \langle\Delta(p),\hat{q}\rangle-\langle p,\Delta^*(\hat{q})\rangle=\rmd_h\omega \end{equation} for all~$p\in P$, $\hat{q}\in\hat{Q}$ and some~$\omega=\omega(p,\hat{q})\in\La_h^{n-1}(\pi)$, where~$\langle\cdot\,,\cdot\rangle$ is the natural pairing between the module and its adjoint. It is useful to keep in mind that the correspondence~$(p,\hat{q})\mapsto\omega(p,\hat{q})$ is a $\mathcal{C}$-differential operator with respect to both arguments. An operator~$\Delta$ is called \emph{self-adjoint} if~$\Delta^*=\Delta$ and \emph{skew-adjoint} if~$\Delta^*=-\Delta$. \begin{coordinates} \label{sec:gjsis_jets:vari-compl-lagr-1} If~$\Delta=\sum_I a_ID_i$ is a scalar operator then \begin{equation}\label{eq:gjsis_jets:31} \Delta^*=\sum_I(-1)^{\abs{I}}D_I\circ a_I. \end{equation} For a matrix $\mathcal{C}$-differential operator~$\Delta=\left\Vert\Delta_{ij}\right\Vert$ one has \begin{equation}\label{eq:gjsis_jets:32} \Delta^*=\left\Vert\Delta_{ji}^*\right\Vert, \end{equation} where~$\Delta_{ji}^*$ is given by \eqref{eq:gjsis_jets:31}. \end{coordinates} We can now describe the groups~$E_1^{p,q}(\pi)$. \begin{theorem}[One-line Theorem] \label{thm:gjsis_jets:2} Let~$\pi\colon E\to M$, $\dim M=n$, be a locally trivial vector bundle. Then: \begin{enumerate} \item the groups~$E_1^{0,q}(\pi)$, $q=0,\dots,n-1$, are isomorphic to the de~Rham cohomology groups~$H^q(M)$ of the manifold~$M$; \item the group~$E_1^{0,n}(\pi)$ consists of the Lagrangians depending on the fields that are sections of~$\pi$; \item the groups~$E_1^{p,n}(\pi)$, $p>0$, are identified with the modules~$\CDiff^{\mathrm{sk\,*}}_{p-1}(\kappa(\pi),\hat{\kappa}(\pi))$ of $(p-1)$-linear skew-symmetric $\mathcal{C}$-differential operators that are skew-adjoint in each argument (in particular, $E_1^{1,n}(\pi)=\hat{\kappa}(\pi)$); \item all other terms are trivial. \end{enumerate} \end{theorem} \begin{remark} The group~$E_1^{0,n}(\pi)$ is also called the $n$th horizontal cohomology group of~$\pi$ and denoted by~$H_h^n(\pi)$. \end{remark} \begin{remark} \label{rem:gjsis_jets:3} The construction above is a particular case of Vinogradov's \emph{$\mathcal{C}$-spectral sequence},~\cite{Vinogradov:AlGFLFT,Vinogradov:SSAsNDEqAlGFLFTC,Vinogradov:SSLFCLLTNT,Vinogradov:CAnPDEqSC,KrasilshchikVerbovetsky:HMEqMP}. \end{remark} In what follows, we shall assume the manifold~$M$ to be cohomologically trivial, i.e., its de~Rham cohomology is isomorphic to~$\mathbb{R}$. Define the operator~$\delta\colon\La_h^n(\pi)\to\hat{\kappa}(\pi)$ as the composition of the projection~$\La_h^n(\pi)\to E_1^{0,n}$ and the differential~$\delta_1^{0,n}\colon E_1^{0,n}\to E_1^{1,n}$. \begin{remark} In what follows, we shall also use the notation~$\delta$ for the differential~$\delta_1^{0,n}$ itself. \end{remark} \begin{coordinates} If~$\omega\in\La_h^n(\pi)$ is of the form~$L\mathinner{\!}\mathrm{d} x^1\wedge\dots\wedge\mathinner{\!}\mathrm{d} x^n$ then \begin{equation}\label{eq:gjsis_jets:33} \delta(\omega)= \left(\frac{\delta L}{\delta u^1},\dots,\frac{\delta L}{\delta u^m}\right), \end{equation} where \begin{equation}\label{eq:gjsis_jets:34} \frac{\delta L}{\delta u^j}=\sum_I(-1)^{\abs{I}}D_I\frac{\partial L}{\partial u_I^j} \end{equation} are \emph{variational derivatives}. Thus,~$\delta$ is the \emph{Euler} operator and it takes a Lagrangian density~$\omega$ to the corresponding \emph{Euler operator}. \end{coordinates} \begin{proposition} \label{prop:gjsis_jets:4} Let~$\pi$ be a locally trivial vector bundle over a cohomologically trivial manifold~$M$. Then the complex \begin{multline}\label{eq:gjsis_jets:35} \mathcal{F}(\pi)\xrightarrow{\rmd_h^{0,0}}\La_h^1(\pi)\xrightarrow{\rmd_h^{0,1}} \cdots\xrightarrow{}\La_h^{n-1}(\pi)\xrightarrow{\rmd_h^{0,n-1}}\La_h^n(\pi) \\ \xrightarrow{\delta}\hat{\kappa}\xrightarrow{\delta_1^{1,n}} \CDiff^{\mathrm{sk\,*}}_1(\kappa,\hat{\kappa})\xrightarrow{\delta_1^{2,n}} \CDiff^{\mathrm{sk\,*}}_2(\kappa,\hat{\kappa})\xrightarrow{}\cdots \end{multline} is exact, i.e., the kernel of each differential coincides with the image of the preceding one. In the coordinate-free way, the Euler-Lagrange operator can be computed by \begin{equation}\label{eq:gjsis_jets:36} \delta(\omega)=\ell_\omega^*(1),\qquad\omega\in\La_h^n(\pi), \end{equation} while the differentials~$\delta_1^{p,n}$ enjoy the equality \begin{equation}\label{eq:gjsis_jets:37} (\delta_1^{p,n}\Delta)(\phi_1,\dots,\phi_p)= \sum_{i=1}^p(-1)^{i+1} \ell_{\Delta,\phi_1,\dots,\hat{\phi}_i,\dots,\phi_p}(\phi_i) +(-1)^p\ell_{\Delta,\phi_1,\dots,\phi_{p-1}}^*(\phi_p), \end{equation} where~$\Delta\in\CDiff^{\mathrm{sk\,*}}_{p-1}(\kappa(\pi),\hat{\kappa}(\pi))$ and~$\phi_1,\dots,\phi_p\in\kappa$. Here and below we use the notation~$\ell_{\Delta,\phi_1,\dots,\phi_k}(\phi)=\bi{E}_\phi(\Delta)(\phi_1,\dots,\phi_k)$ In particular, if~$\psi\in\hat{\kappa}(\pi)$ then \begin{equation}\label{eq:gjsis_jets:38} \delta_1^{1,n}(\psi)=\ell_\psi-\ell_\psi^*. \end{equation} \end{proposition} \begin{remark} Complex~\eqref{eq:gjsis_jets:35} is exact starting from the term~$\La_h^n(\pi)$ independently of cohomological properties of the manifold~$M$. This complex is called the \emph{global variational complex} of the bundle~$\pi$, see~\cite{Vitolo:VS}. \end{remark} As a consequence of Proposition~\ref{prop:gjsis_jets:4}, we obtain the following result: \begin{theorem} \label{thm:gjsis_jets:3} For a vector bundle~$\pi$ one has: \begin{enumerate} \item The action functional \begin{equation*} s\mapsto\int_Mj_\infty(s)^*(\omega),\qquad s\in\Gamma(\pi),\quad \omega\in\La_h^n(\pi), \end{equation*} is stationary on a section~$s$ if and only if~$j_\infty(s)^*(\delta(\omega))=0$ (i.e., as we shall see below,~$s$ is a solution of the Euler-Lagrange equation corresponding to~$\omega$). \item Variationally trivial Lagrangians are total divergences, which amounts to the equality~$\ker\delta=\im\rmd_h^{0,n-1}$. \item All null total divergences are total curls, i.e., $\rmd_h^{0,n-1}\omega=0$ if and only if~$\omega=\rmd_h^{0,n-2}\theta$ for some~$\theta\in\La_h^{n-2}(\pi)$. \item If~$\psi\in\hat{\kappa}(\pi)$ then the nonlinear differential operator~$\Delta_\psi\colon\kappa(\pi)\to\hat{\kappa}(\pi)$ is of the form~$\delta\omega$ (i.e., is an Euler-Lagrange operator) if and only if~$\ell_\psi=\ell_\psi^*$ (the Helmholtz condition). \end{enumerate} \end{theorem} \subsection{A parallel with finite-dimensional differential geometry. I} \label{sec:gjsis_jets:parallel-with-finite} We want to indicate here a very useful and productive analogy between the geometry of jet spaces (and, more generally, of differential equations) and classical differential geometry of finite-dimensional smooth manifolds. This parallel was exposed by A.M.~Vinogradov first within his philosophy of \emph{Secondary Calculus} (cf.~\cite{Vinogradov:CAnPDEqSC} and references therein) and we elaborate it further. Two points of view on jet spaces as geometrical objects may exist. The first one is formal, traditional and straightforward. It was described on the previous pages and treats~$J^\infty(\pi)$ as a particular case of general infinite-dimensional manifolds. Such an approach, being by all means necessary for a rigorous exposition of the theory, actually ignores essential intrinsic structures of jet spaces actually. Another viewpoint is completely informal but incorporates these structures \emph{ab ovo} and allows to reveal new and non-trivial relations and results just ``translating'' from the language of the classical differential geometry. To facilitate such translations, we shall compile a sort of a dictionary. So, we consider the space~$J^\infty(\pi)$ endowed with the Cartan distribution~$\mathcal{C}$ and take for the points of the new ``manifold'' maximal integral submanifolds of~$\mathcal{C}$. As it was indicated above, they are graphs of~$j_\infty(s)$, $s\in\Gamma(\pi)$, and thus new ``points'' are sections of~$\pi$ (i.e., fields). Let~$\omega\in\La_h^n(\pi)$ be a horizontal $n$-form on~$J^\infty(\pi)$ (or a Lagrangian density). Then to any ``point''~$s\in\Gamma(\pi)$ we can put into correspondence the number \begin{equation*} \omega(s)=\int_M j_\infty(s)^*(\omega). \end{equation*} Thus, Lagrangians are understood as functions. Due to the identity \begin{equation*} j_\infty(s)^*(\rmd_h\omega)=\mathinner{\!}\mathrm{d}(j_\infty(s)^*(\omega)) \end{equation*} and the Stokes formula \begin{equation*} \int_M\mathinner{\!}\mathrm{d}\theta=\int_{\partial M}\theta,\qquad\theta\in\Lambda^{n-1}(M), \end{equation*} ``functions'' of the form~$\rmd_h\omega$ vanish at all points. So, no-trivial functions are elements of the cohomology group~$E_1^{0,n}(\pi)=H_h^h(\pi)$. Thus, the beginning of the dictionary is \begin{eqnarray*} \text{\textbf{Manifold~$M$}}&&\text{\textbf{Jet space~$J^\infty(\pi)$}} \\[1ex] \text{points}&\quad\longleftrightarrow\quad &\text{sections of~$\pi$ (fields)} \\ \text{functions}&\quad\longleftrightarrow\quad&\text{Lagrangians }\ \omega=L\mathinner{\!}\mathrm{d} x^1\wedge\dots\wedge\mathinner{\!}\mathrm{d} x^n\\ \text{value at a point, }\ f(x)&\quad\longleftrightarrow\quad& \text{integral }\ \omega(s)=\int_Mj_\infty(s)^*\omega \\ &&\text{(the cohomology class of~$\omega$)} \end{eqnarray*} The next step is to define vector fields. They should be infinitesimal transformations of~$J^\infty(\pi)$ that preserve the Cartan distribution (or, equivalently, move ``points'' to ``points''). These are exactly the fields lying in~$\X_{\scriptscriptstyle{\mathcal{C}}}(\pi)$. But vector fields~$X\in\mathcal{C}\mathcal{X}(\pi)$ (i.e., lying in the Cartan distribution) are tangent to maximal integral manifolds of the latter and thus are trivial in the space of fields. Consequently, non-trivial vector fields are identified with elements of~$\sym\pi$, i.e., with evolutionary vector fields on~$J^\infty(\pi)$. On the other hand, as it was indicated above, they are integrable sections\footnote{By integrable sections we mean those ones whose graphs are maximal integral manifolds of the Cartan distribution.} of~$J_h^\infty(\kappa)$. Hence, this bundle can be considered as the tangent bundle to~$J^\infty(\pi)$. So, the dictionary can be continued as follows: \begin{eqnarray*} \text{\textbf{Manifold~$M$}}&&\text{\textbf{Jet space~$J^\infty(\pi)$}} \\[1ex] \text{vector fields, $\mathcal{X}(M)$}&\quad\longleftrightarrow\quad& \text{evolutionary vector fields, $\kappa(\pi)$} \\ \text{the tangent bundle}&\quad\longleftrightarrow\quad& \text{the bundle of horizontal jets $J_h^\infty(\kappa(\pi))$} \end{eqnarray*} Differential forms on a smooth manifold~$M$ may be understood as multi-linear functions on the space of vector fields (or fibre-wise multi-linear functions on~$T(M)$): we insert a vector field into a $p$-form and obtain a $(p-1)$-form. In our context, such objects are exactly elements of~$\CDiff^{\mathrm{sk\,*}}_{p-1}(\kappa,\hat{\kappa})=E_1^{p,n}(\pi)$. We call them \emph{variational forms} of degree~$p$ and have the following parallel: \begin{eqnarray*} \text{\textbf{Manifold~$M$}}&&\text{\textbf{Jet space~$J^\infty(\pi)$}} \\[1ex] \text{differential forms, $\Lambda^p(M)$}&\quad\longleftrightarrow\quad&\text{variational forms, $\CDiff^{\mathrm{sk\,*}}_{p-1}(\kappa,\hat{\kappa})$} \\ \text{the de~Rham complex}&\quad\longleftrightarrow\quad&\text{the variational complex} \end{eqnarray*} \begin{remark} \label{rem:gjsis_jets:4} It can also be shown that smooth maps~$J^\infty(\pi)\toJ^\infty(\pi')$ preserving the Cartan distributions are completely determined by non-linear differential operators from~$\Gamma(\pi)$ to~$\Gamma(\pi')$ while the differentials of these maps~$J_h^\infty(\kappa)\toJ_h^\infty(\kappa')$ correspond to linearizations. Unfortunately, a detailed exposition of this parallel is out of scope of our review. \end{remark} We shall continue to compile our dictionary in Subsection~\ref{sec:gjsis_jets:parallel-with-finite-1}. \subsection{Hamiltonian formalism} \label{sec:gjsis_jets:hamilt-form} The objects dual to~$\CDiff^{\mathrm{sk\,*}}_{p-1}(\kappa(\pi),\hat{\kappa}(\pi))$ are the modules of \emph{variational multivectors}~$D_p(\pi)=\CDiff^{\mathrm{sk\,*}}_{p-1}(\hat{\kappa}(\pi),\kappa(\pi))$. In particular, $D_1(\pi)=\kappa(\pi)$. We also set~$D_0(\pi)=\La_h^n(\pi)/\im \rmd_h^{0,n-1}=E_1^{0,n}(\pi)$. To describe Hamiltonian formalism on~$J^\infty(\pi)$, we first introduce the \emph{variational Schouten bracket}~\cite{KerstenKrasilshchikVerbovetsky:HOpC} \begin{equation*} [\![\cdot,\cdot]\!]\colon D_p(\pi)\times D_q(\pi)\to D_{p+q-1}(\pi) \end{equation*} in the following way (cf.~\cite{Krasilshchik:SBCAl,IgoninVerbovetskyVitolo:VMBGJS}, see also~\cite{Dorfman:DSInNEvEq,GelfandDorfman:SBHOp}). If~$B=[\omega]\in D_0(\pi)$ is a coset of a horizontal form~$\omega\in\La_h^n(\pi)$ and~$A\in D_p(\pi)$, $p>0$, then we set \begin{equation*} [\![ A,B]\!]=(-1)^p[\![ B,A]\!]=A(\delta\omega), \end{equation*} while for any~$B\in D_q(\pi)$, $q>0$, and~$\psi=\delta\omega\in\hat{\kappa}(\pi)$, $\omega\in\La_h^n(\pi)$, \begin{equation*} [\![ A,B]\!](\psi)=[\![ A,B(\psi)]\!]+(-1)^{q-1}[\![ A(\psi),B]\!], \end{equation*} and these two equalities define the bracket completely. In particular, \begin{equation*} [\![ \phi,[\omega]]\!]=[\bi{E}_\phi(\omega)],\qquad\phi\in D_1(\pi)=\kappa(\pi),\quad \omega\in\La_h^n(\pi), \end{equation*} and \begin{equation*} [\![ \phi,\phi']\!]=\bi{E}_\phi(\phi')-\bi{E}_{\phi'}(\phi)=\{\phi,\phi'\},\qquad \phi,\phi'\in D_1(\pi). \end{equation*} \begin{proposition} \label{prop:gjsis_jets:5} The variational Schouten bracket determines a super Lie algebra structure in the space~$D(\pi)=\sum_{p\ge 0}D_p(\pi)$ in the following sense: \begin{align} \label{eq:gjsis_jets:39} &[\![ A,B]\!]=-(-1)^{(p-1)(q-1)}[\![ B,A]\!],\\\label{eq:gjsis_jets:40} &(-1)^{(p+1)(r+1)}[\![\lshad A,B]\!],C]\!]+(-1)^{(q+1)(p+1)}[\![\lshad B,C]\!],A]\!]\\ &\qquad+(-1)^{(r+1)(q+1)}[\![\lshad C,A]\!],B]\!]=0\nonumber \end{align} for all~$A\in D_p(\pi)$, $B\in D_q(\pi)$, $C\in D_r(\pi)$. \end{proposition} To compute the Schouten bracket explicitly, for any natural~$n$ consider the set~$S_n^i$ of all~$(n-i)$-\emph{un-shuffles} consisting of all permutations~$\sigma$ of the set~$\{1,\dots,n\}$ such that \begin{equation*} \sigma(1)<\dots<\sigma(i),\qquad\sigma(i+1)<\dots<\sigma(n). \end{equation*} We formally set~$S_n^i=\varnothing$ for~$i<0$ and~$i>n$. We also use a short notation~$\psi_{\sigma(k_1,k_2)}$ for~$\psi_{\sigma(k_1)},\dots,\psi_{\sigma(k_2)}$. Let now~$A\in D_p(\pi)$ and~$B\in D_q(\pi)$. Then for any~$\psi_1,\dots,\psi_{p+q-1}\in\hat{\kappa}(\pi)$ we have \begin{multline}\label{eq:gjsis_jets:41} [\![ A,B]\!](\psi_1,\dots,\psi_{p+q-1}) =\sum_{\sigma\in S_{p+q-1}^{q-1}}(-1)^\sigma \ell_{B,\psi_{\sigma(1,q-1)}}(A(\psi_{\sigma(q,p+q-1)})) \\ -(-1)^{(p-1)(q-1)}\sum_{\sigma\in S_{p+q-1}^p}(-1)^\sigma B(\ell^*_{A,\psi_{\sigma(1,p-1)}}(\psi_{\sigma(p)}), \psi_{\sigma(p+1,p+q-1)}) \\ -(-1)^{(p-1)(q-1)}\sum_{\sigma\in S_{p+q-1}^{p-1}}(-1)^\sigma \ell_{A,\psi_{\sigma(1,p-1)}}(B(\psi_{\sigma(p,p+q-1)})) \\ +\sum_{\sigma\in S_{p+q-1}^q}(-1)^\sigma A(\ell^*_{B,\psi_{\sigma(1,q-1)}} (\psi_{\sigma(q)}),\psi_{\sigma(q+1,p+q-1)}), \end{multline} where~$(-1)^\sigma$ stands for the parity of the permutation~$\sigma$ and, as before,~$\ell_{\Delta,\psi_1,\dots,\psi_k}(\phi)=\bi{E}_\phi(\Delta)(\psi_1,\dots,\psi_k)$. We say that a bivector~$A\in D_2(\pi)=\CDiff^{\mathrm{sk\,*}}(\hat{\kappa}(\pi),\kappa(\pi))$ is a \emph{Hamiltonian structure} on~$J^\infty(\pi)$ if \begin{equation}\label{eq:gjsis_jets:42} [\![ A,A]\!]=0. \end{equation} \begin{remark} A more appropriate name is a \emph{Poisson} structure but we follow here the tradition accepted in the theory of integrable systems. \end{remark} Given a Hamiltonian structure, one can define a \emph{Poisson bracket} (with respect to the Hamiltonian structure~$A$) on the set of Lagrangians: \begin{equation}\label{eq:gjsis_jets:43} \{\omega,\omega'\}_A=\langle A(\delta(\omega)),\delta(\omega')\rangle,\qquad \omega,\omega'\in D_0(\pi). \end{equation} Two elements are \emph{in involution} (with respect to the structure~$A$) if \begin{equation*} \{\omega,\omega'\}_A=0. \end{equation*} \begin{proposition} \label{prop:gjsis_jets:6} For any~$A\in\CDiff^{\mathrm{sk\,*}}(\hat{\kappa}(\pi),\kappa(\pi))$ one has \begin{equation}\label{eq:gjsis_jets:44} \{\omega,\omega'\}_A=-\{\omega',\omega\}_A \end{equation} If in addition~$A$ satisfies~\eqref{eq:gjsis_jets:42} then \begin{equation}\label{eq:gjsis_jets:45} \{\omega,\{\omega',\omega''\}_A\}_A+ \{\omega',\{\omega'',\omega\}_A\}_A+ \{\omega'',\{\omega,\omega'\}_A\}_A=0. \end{equation} One also has \begin{equation}\label{eq:gjsis_jets:46} A_{\{\omega,\omega'\}_A}=\{A_\omega,A_{\omega'}\}, \end{equation} where~$A_\omega=A(\delta(\omega))\in\kappa(\pi)$ and the curlies in the right-hand side denote the Jacobi bracket. \end{proposition} Chose a Hamiltonian structure~$A$ and consider the sequence of operators \begin{equation}\label{eq:gjsis_jets:47} \hspace*{-6pc}\qquad 0\xrightarrow{}D_0(\pi)\xrightarrow{\partial_A}D_1(\pi)\xrightarrow{} \cdots\xrightarrow{}D_p(\pi)\xrightarrow{\partial_A}D_{p+1}(\pi) \xrightarrow{}\cdots, \end{equation} where~$\partial_A(B)=[\![ A,B]\!]$. \begin{proposition} \label{prop:gjsis_jets:7} Sequence~\eqref{eq:gjsis_jets:47} is a complex, i.e.,~$\partial_A\circ\partial_A=0$. \end{proposition} We say that~$\bi{E}_\phi$ is a \emph{Hamiltonian vector field} if~$\phi\in\ker\partial_A$. This is equivalent to \begin{equation}\label{eq:gjsis_jets:48} \bi{E}_\phi(\{\omega,\omega'\}_A)=\{\bi{E}_\phi(\omega),\omega'\}_A +\{\omega,\bi{E}_\phi(\omega')\}_A, \end{equation} i.e., $\bi{E}_\phi$ preserves the Poisson bracket. Due to Proposition~\ref{prop:gjsis_jets:7}, a particular case of Hamiltonian fields are fields of the form~$\bi{E}_{A(\delta(\omega))}$. In this case,~$\omega\in D_0(\pi)$ is called the \emph{Hamiltonian} of the field under consideration. We say that~$\omega\in D_0(\pi)$ is a \emph{first integral} of a Hamiltonian field~$\bi{E}_\phi$ if~$\bi{E}_\phi(\omega)=0$. A Hamiltonian field~$\bi{E}_{\phi'}$ is a \emph{symmetry} for the field~$\bi{E}_\phi$ if~$[\bi{E}_{\phi'},\bi{E}_\phi]=0$, or \begin{equation*} (\bi{E}_\phi-\ell_\phi)(\phi')=0. \end{equation*} \begin{proposition} \label{prop:gjsis_jets:8} If~$\bi{E}_\phi$ is a Hamiltonian vector field with respect to a Hamiltonian structure~$A$ then the operator~$A\circ\delta=\partial_A\colon D_0(\pi)\to D_1(\pi)$ takes first integrals of~$\bi{E}_\phi$ to its symmetries. \end{proposition} A Hamiltonian structure~$B\in D_2(\pi)$ is said to be \emph{compatible} with the structure~$A$ if~$B\in\ker\partial_A$, or \begin{equation}\label{eq:gjsis_jets:49} [\![ A,B]\!]=0. \end{equation} This is equivalent to the fact that all the bivectors \begin{equation}\label{eq:gjsis_jets:50} \lambda A+\mu B,\qquad \lambda,\mu\in\mathbb{R}, \end{equation} are Hamiltonian structures on~$J^\infty(\pi)$. The family~\eqref{eq:gjsis_jets:50} is called a \emph{Poisson pencil}. When two Hamiltonian structures are given, one also says that they form a \emph{bi-Hamiltonian structure}. \begin{coordinates} Let us indicate how to verify conditions~\eqref{eq:gjsis_jets:42} and~\eqref{eq:gjsis_jets:49} in coordinates (the explanation will be given below in Subsection~\ref{sec:gjsis_jets:parallel-with-finite-1}, see Remark~\ref{rem:gjsis_jets:5}). Take bivectors~$A$, $B\in D_2(\pi)=\CDiff^{\mathrm{sk\,*}}_1(\hat{\kappa}(\pi),\kappa(\pi))$. Then~$A$ and~$B$, in adapted coordinates in~$J^\infty(\pi)$, are represented as matrix $\mathcal{C}$-differential operators \begin{equation*} A=\left\Vert\sum_\sigma a_\sigma^{ij}D_\sigma\right\Vert,\qquad B=\left\Vert\sum_\sigma b_\sigma^{ij}D_\sigma\right\Vert, \end{equation*} where~$i$, $j=1,\dots, m=\dim\pi$. Let us put into correspondence to these operators the functions \begin{equation}\label{eq:gjsis_jets:58} W_A=\sum_{\sigma,i,j} a_\sigma^{ij}p_\sigma^ip^j,\qquad W_B=\sum_{\sigma,i,j} b_\sigma^{ij}p_\sigma^ip^j, \end{equation} where~$p_\sigma^i$ are \emph{odd} variables. Then~$A$ is a Hamiltonian structure if and only if \begin{equation}\label{eq:gjsis_jets:51} \delta\left(\sum_i\frac{\delta W_A}{\delta u^i}\frac{\delta W_A}{\delta p^i}\right)=0, \end{equation} while two Hamiltonian structures are compatible if and only if \begin{equation}\label{eq:gjsis_jets:52} \delta\left(\sum_i\left(\frac{\delta W_A}{\delta u^i}\frac{\delta W_B}{\delta p^i}+\frac{\delta W_B}{\delta u^i}\frac{\delta W_A}{\delta p^i}\right)\right)=0. \end{equation} \end{coordinates} \begin{theorem}[the Magri Scheme, see~\cite{Magri:SMInHEq,Krasilshchik:SBCAl}] \label{thm:gjsis_jets:4} Let~$(A,B)$ be a bi-Hamiltonian structure on~$J^\infty(\pi)$ and assume that the complex~\eqref{eq:gjsis_jets:47} is acyclic in the term~$D_1(\pi)$, i.e., every Hamiltonian vector field with respect to~$A$ possesses a Hamiltonian. Assume also that two densities $\omega_1$, $\omega_2\in D_0(\pi)$ are given, such that $\partial_{A}(\omega_1)=\partial_{B}(\omega_2)$. Then: \begin{enumerate} \item There exist elements $\omega_3,\dots,\omega_s,\dots\in D_0(\pi)$ satisfying \begin{equation}\label{eq:gjsis_jets:60} \partial_{A}(\omega_s)=\partial_{B}(\omega_{s+1}),\qquad s=2,3,\dots \end{equation} \item All elements $\omega_1,\dots,\omega_s,\dots$ are in involution with respect to both Hamiltonian structures, i.e., \begin{equation*} \{\omega_\alpha,\omega_\beta\}_A=\{\omega_\alpha,\omega_\beta\}_B=0 \end{equation*} for all $\alpha$, $\beta\ge1$. \end{enumerate} \end{theorem} \begin{example}[the KdV hierarchy] \label{exmp:gjsis_jets:1} Consider~$J^\infty(\pi)$ for the trivial one-dimensional bundle~$\pi\colon\mathbb{R}\times\mathbb{R}\to\mathbb{R}$. Let~$x$ be the independent variable and~$u$ be the fibre coordinate (the unknown function). Then the adapted coordinates~$u=u_0$, $u_1,\dots,u_k,\dots$ in~$J^\infty(\pi)$ arise, where~$u_k$ corresponds to~$\partial^ku/\partial x^k$. The operators \begin{equation*} A=D_x=\frac{\partial}{\partial x} +\sum_{k=0}^\infty u_{k+1}\frac{\partial}{\partial u_k} \end{equation*} and \begin{equation*} B=D_x^3+4uD_x+2u_1, \end{equation*} as it can be easily checked using~\eqref{eq:gjsis_jets:51} and~\eqref{eq:gjsis_jets:52}, constitute a bi-Hamiltonian structure on~$J^\infty(\pi)$. Then obviously for the horizontal forms \begin{equation*} \omega_1=\frac{1}{2}u^2\mathinner{\!}\mathrm{d} x,\qquad\omega_2=\frac{1}{2}u\mathinner{\!}\mathrm{d} x \end{equation*} one has~$\partial_A\omega_1=A(u)=u_1$ and~$\partial_B\omega_2=B(1)=u_1$, i.e., \begin{equation*} \partial_A\omega_1=\partial_B\omega_2. \end{equation*} The first cohomology group of~$\partial_A$ is trivial (see~\cite{Getzler:DTHOpFCV}), and consequently we obtain an infinite series of first integrals and the corresponding symmetries. The second, after~$u_1$, symmetry is~$6uu_1+u_3$: \begin{equation*} 6uu_1+u_3=\partial_A((u^3-\frac{1}{2}u_1^2)\mathinner{\!}\mathrm{d} x)=\partial_B(\frac{1}{2}u^2\mathinner{\!}\mathrm{d} x). \end{equation*} The corresponding flow on~$J^\infty(\pi)$ is governed by the evolution equation \begin{equation*} u_t=6uu_x+u_{xxx}; \end{equation*} thus, we obtain the Korteweg-de Vries equation and the corresponding hierarchy of commuting flows (the higher KdV equations). The entire family of commuting flows can be obtained by applying the \emph{Lenard recursion operator} (see~\cite{GardnerGreeneKruskalMiura:KEqGVMExS}) \begin{equation}\label{eq:gjsis_jets:53} R=B\circ A^{-1}=D_x^2+4u+2u_1D_x^{-1} \end{equation} to the right-hand side of the first flow~$u_t=u_x$ sufficiently many times. \end{example} \begin{example}[the Boussinesq hierarchy] \label{exmp:gjsis_jets:2} Consider the adapted coordinates~$x$, $u$, $v,\dots,u_k$, $v_k,\dots$ in the space~$J^\infty(\pi)$, where~$\pi\colon\mathbb{R}^2\times\mathbb{R}\to\mathbb{R}$ is the trivial two-dimensional bundle over~$\mathbb{R}$. Then the operators \begin{equation*} A= \begin{pmatrix} 0&D_x\\ D_x&0 \end{pmatrix},\qquad B= \begin{pmatrix} \sigma D_x^3+uD_x+\frac{1}{2}u_1&\frac{1}{2}vD_x\\ \frac{1}{2}vD_x+\frac{1}{2}v_1&D_x \end{pmatrix}, \end{equation*} where~$\sigma$ is a real constant, form a bi-Hamiltonian structure. For the differential forms~$\omega_1=2(u+v)\mathinner{\!}\mathrm{d} x$ and~$\omega_2=uv\mathinner{\!}\mathrm{d} x$ one obviously has~$\partial_A\omega_1=\partial_B\omega_2$. The arising hierarchy of commuting flows corresponds to the evolution equation \begin{equation}\label{eq:gjsis_jets:54} \begin{array}{l} u_t =u_xv+uv_x+\sigma v_{xxx},\\ v_t =u_x+vv_x \end{array} \end{equation} which is the \emph{two-component Boussinesq system} which can be obtained from the Kaup equation, see~\cite{Kaup:HOrWWEqMSIt}. Note that there exists another Hamiltonian operator \begin{equation*} C= \begin{pmatrix} C^{uu}&C^{uv}\\ C^{vu}&C^{vv} \end{pmatrix}, \end{equation*} where \begin{align*} C^{uu}&=\sigma D_x^3+\frac{3}{2}\sigma v_1D_x^2+(\sigma v_2+uv)D_x+\frac{1}{2}(\sigma v_3+uv_1+u_1v),\\ C^{uv}&=\sigma D_x^3+(u+\frac{1}{4}v^2)D_x+\frac{1}{2}u_1,\\ C^{vu}&=\sigma D_x^3+(u+\frac{1}{4}v^2)D_x+\frac{1}{2}(u_1+vv_1),\\ C^{vv}&=vD_x+\frac{1}{4}v_1, \end{align*} which is compatible both with~$A$ and~$B$. In this sense, system~\eqref{eq:gjsis_jets:54} is \emph{tri-Hamiltonian}. \end{example} \begin{example}[the KdV hierarchy, II] \label{exmp:gjsis_jets:3} Let~$\pi\colon\mathbb{R}^3\times\mathbb{R}^1\to\mathbb{R}^1$ be the trivial three-dimensional bundle with the coordinates~$t$ in the base and~$u$, $v$, $w$ in the fibre. Introduce the adapted coordinates~$u_k$, $v_k$, $w_k$, where $k=0,1,2,\dots$, in~$J^\infty(\pi)$ and consider the operators \begin{equation*} A= \begin{pmatrix} 0&-1&0\\ 1&0&-6u\\ 0&6u&D_t \end{pmatrix},\quad B= \begin{pmatrix} 0&-2u&-D_t-2v\\ 2u&D_t&-12u^2-2w\\ -D_t+2v&12u^2+2w&8uD_t+4u_1 \end{pmatrix}, \end{equation*} which form a bi-Hamiltonian structure on~$J^\infty(\pi)$. It can be easily seen that~$\partial_A\omega_1=\partial_B\omega_2$, where \begin{equation*} \omega_1=(uw-\frac{1}{2}v^2+2u^3)\mathinner{\!}\mathrm{d} t,\qquad \omega_2=-(\frac{3}{2}u^2+\frac{1}{2}w)\mathinner{\!}\mathrm{d} t. \end{equation*} Thus, we obtain a hierarchy of commuting flows whose first term is \begin{equation*} u_x=v,\qquad v_x=w,\qquad w_x=u_t-6uv, \end{equation*} which is obviously the KdV equation rewritten in a different way (cf.~the paper~\cite{Tsarev:HPSIEqCMMMP}). Note that the Lenard recursion operator~\eqref{eq:gjsis_jets:53} in the new representation of the KdV hierarchy acquires the form \begin{equation*} R= \begin{pmatrix} 0&-2u&-D_t-2v\\ 2u&D_t&-12u^2-2w\\ -D_t+2v&12u^2-2w&8uD_t+4u_1 \end{pmatrix} \circ \begin{pmatrix} -36uD_t^{-1}\circ u&1&-6uD_t^{-1}\\ -1&0&0\\ 6D_t^{-1}\circ u&0&D_t^{-1} \end{pmatrix}. \end{equation*} \end{example} \begin{remark} In a recently published paper~\cite{BarakatSoleKac:PVAlTHEq}, the authors formulate a much weaker than triviality of the first Poisson cohomology group criterion for feasibility of the Magri scheme. The criterion is given in the framework of Dirac structures~\cite{Dorfman:DSInNEvEq,Courant:DM} (though it also admits a self-contained formulation) that unfortunately lie beyond the scope of our review. Note that Dirac structures that merge the notions of symplectic and Hamiltonian operators constitute an interesting object for geometrical research in PDEs. \end{remark} \begin{remark} Normal forms for Hamiltonian operators of order~$\le 5$ and a ``variational'' analog of the Darboux Lemma were presented in~\cite{Astashov:NFHOpFT,AstashovVinogradov:SHOpFT}. In~\cite{SoleKacWakimoto:CPVAl}, one can find normal forms for operators of order~$\le 11$ and some classification results for operators of higher order. \end{remark} \subsection{A parallel with finite-dimensional differential geometry. II} \label{sec:gjsis_jets:parallel-with-finite-1} Let us now continue to construct the dictionary started in Section~\ref{sec:gjsis_jets:parallel-with-finite}. Of course, variational multivectors introduced in Section~\ref{sec:gjsis_jets:hamilt-form} are exact counterparts of classical multivector fields in differential geometry. These fields are naturally understood as smooth functions on~$T^*M$ skew-symmetric and multi-linear with respect to fibre variables. Exactly the same interpretation is valid for variational vectors if one considers the bundle~$\tau^*\colonJ_h^\infty(\hat{\kappa})\toJ^\infty(\pi)$\footnote{Note that in such a way we independently arrived to Kupershmidt's notion of the cotangent bundle to a vector bundle, see~\cite{Kupershmidt:GJBSLHF}.}. Thus, we have the following translations: \begin{eqnarray*} \text{\textbf{Manifold~$M$}}&&\text{\textbf{Jet space~$J^\infty(\pi)$}} \\[1ex] \text{multivector fields}&\quad\longleftrightarrow\quad&\text{variational multivectors, $\CDiff^{\mathrm{sk\,*}}_{p-1}(\hat{\kappa},\kappa)$} \\ \text{Schouten bracket}&\quad\longleftrightarrow\quad&\text{variational Schouten bracket} \\ \text{Poisson structure}&\quad\longleftrightarrow\quad&\text{Hamiltonian operator} \\ \text{cotangent bundle,}&\quad\longleftrightarrow\quad&\text{variational cotangent bundle,}\\ \qquad\qquad T^*M\to M&\quad\longleftrightarrow\quad J_h^\infty(\hat{\kappa})\toJ^\infty(\pi)\end{eqnarray*} As it is well known, the tangent space~$T^*M$ is endowed with the natural symplectic structure~$\Omega=\mathinner{\!}\mathrm{d} p\wedge\mathinner{\!}\mathrm{d} q\in\Lambda^2(M)$ and, in addition,~$\Omega=\mathinner{\!}\mathrm{d}\rho$, where the form~$\rho=p\mathinner{\!}\mathrm{d} q$ is defined invariantly as well. Similar constructions exist on~$J_h^\infty(\hat{\kappa})$. Let us show this. To this end, recall (see Remark~\ref{rem:gjsis_jets:1}) that the manifold~$J_h^\infty(\hat{\kappa})$ is diffeomorphic to~$J^\infty(\pi\times_M\hat{\pi})$, where~$\hat{\pi}=\Hom(\pi,\bigwedge^nT^*M)$. Hence, the module of variational $1$-forms on~$J_h^\infty(\hat{\pi})$ is isomorphic to \begin{equation}\label{eq:gjsis_jets:55} \hat{\kappa}(\pi\times_M\hat{\pi})= \hat{\kappa}(\pi)\times_{J^\infty(\pi)}\kappa(\pi). \end{equation} Then the $1$-form~$\rho_\pi$ (the analog of~$p\mathinner{\!}\mathrm{d} q$) is uniquely defined by the condition \begin{equation}\label{eq:gjsis_jets:56} j_h^\infty(\psi)^*(\rho_\pi)=(\psi,0), \end{equation} where~$\psi\in\hat{\kappa}(\pi)$ is an arbitrary variational $1$-form on~$J^\infty(\pi)$. Now, by the same reasons and dually to~\eqref{eq:gjsis_jets:55}, the module of vector fields on~$J_h^\infty(\hat{\kappa})$ is \begin{equation*} \kappa(\pi\times_M\hat{\pi})=\kappa(\pi)\times_{J^\infty(\pi)}\kappa(\hat{\pi}). \end{equation*} Thus we see that the symplectic structure~$\Omega_\pi$ must be an element of the module~$\CDiff^{\mathrm{sk\,*}}(\kappa(\pi\times_M\hat{\pi}),\hat{\kappa}(\pi\times_M\hat{\pi}))$. For any element~$(\phi,\psi)\in\kappa(\pi\times_M\hat{\pi})$ we set \begin{equation}\label{eq:gjsis_jets:57} \Omega_\pi(\phi,\psi)=(-\psi,\phi); \end{equation} this is a skew-adjoint $\mathcal{C}$-differential operator of order~$0$. \begin{remark} The form~$\rho_\pi$ can be defined in a different way. Namely, let~$X=(\phi,\psi)$ be a vector field on~$J_h^\infty(\hat{\kappa})$. Then its 1st component is a vector field vertical with respect to the projection~$J_h^\infty(\hat{\kappa})\toJ^\infty(\pi)$ and may be understood as a function on~$J_h^\infty(\hat{\kappa})$. Then we set~$\rho_\pi(X)=\phi$. This definition is equivalent to~\eqref{eq:gjsis_jets:56}. \end{remark} Of course, the operator~$\Omega_\pi$ is invertible and the inverse one~$S_\pi=\Omega_\pi^{-1}$ is a bivector on the cotangent manifold~$J_h^\infty(\hat{\kappa})$. There exist two points of view at this manifold (as well as at~$T^*M$). The first one treats it as a classical (``even'') manifold. The second approach considers~$J_h^\infty(\hat{\kappa})$ as a super-manifold with odd coordinates along the projection~~$J_h^\infty(\hat{\kappa})\toJ^\infty(\pi)$ and even ones in the base. If one takes the first approach the bivector~$S_\pi$ will define the Poisson bracket for functions on~$J_h^\infty(\hat{\kappa})$. The second approach leads to functions multi-linear and skew-symmetric with respect to fibre variables. As it was stated above, these functions are identified with variational multivectors on~$J^\infty(\pi)$. Then the super-bracket defined by~$S_\pi$ coincides with the Schouten bracket. The bracket is given by the formula \begin{equation} \label{eq:gjsis_jets:59} S_\pi(\delta\omega_1)(\omega_2) =\langle S_\pi(\delta\omega_1),\delta\omega_2\rangle=\left\{ \begin{array}{ll} \{\omega_1,\omega_2\}, &\text{the even case,}\\ [\![\omega_1,\omega_2]\!], &\text{the odd case,} \end{array}\right. \end{equation} in both cases. \begin{remark} \label{rem:gjsis_jets:5} The above said clarifies the meaning of formulas~\eqref{eq:gjsis_jets:58}--\eqref{eq:gjsis_jets:52}. Namely, the correspondence~$A\mapsto W_A$ given by~\eqref{eq:gjsis_jets:58} describes how to construct the function on~$J_h^\infty(\hat{\kappa})$ when a bivector~$A$ is given (to be more precise, this function is the horizontal cohomology class of the form~$W_A\mathinner{\!}\mathrm{d} x^1\wedge\dots\wedge\mathinner{\!}\mathrm{d} x^n$ in~$H_h^n(\pi)$). The argument of~$\delta$ in~\eqref{eq:gjsis_jets:53} is the coordinate expression of the bracket~\eqref{eq:gjsis_jets:59} in the odd case while~\eqref{eq:gjsis_jets:59} itself checks triviality of its horizontal cohomology class. \end{remark} \section{Differential Equations} \label{sec:gjsis_eqs:diff-equat} With the concept of the jet bundle at our disposal we give a geometric definition of differential equations. \subsection{Definition of differential equations} \label{sec:gjsis_eqs:defin-diff-equat} Suppose we have a system \begin{equation} \label{eq:gjsis_eqs:1} F_s(x^i,u^j_I)=0,\qquad s=1,\dots,l, \end{equation} of partial differential equations in $n$ independent variables~$x^i$ and $m$ dependent variables~$u^j$. Equations~\eqref{eq:gjsis_eqs:1} determine a locus in the jet space~$J^{\infty}(\pi)$ of a vector bundle~$\pi\colon E\to M$, such that $\dim E=m+n$, $\dim M=n$. The subset of~$J^{\infty}(\pi)$ defined in this way is not an adequate geometric construction corresponding to the system at hand, because it does not take into account differential consequences of~\eqref{eq:gjsis_eqs:1}. So, we extend~\eqref{eq:gjsis_eqs:1} to a larger system \begin{equation} \label{eq:gjsis_eqs:2} D_I(F_s)=0\qquad\text{for all multi-indices $I$ and $s=1,\dots,l$,} \end{equation} and consider the locus $\mathcal{E}\subsetJ^{\infty}(\pi)$ defined by~\eqref{eq:gjsis_eqs:2}. Thus, we get a correspondence \begin{equation*} F_s(x^i,u^j_I)=0\quad\text{~\eqref{eq:gjsis_eqs:1}}\qquad \mapsto\qquad\mathcal{E}\subsetJ^{\infty}(\pi). \end{equation*} This correspondence behaves nice with respect to solutions of~\eqref{eq:gjsis_eqs:1}: they are those sections of~$\pi$ whose infinite jets lie in~$\mathcal{E}$. To put this another way, the solutions of~\eqref{eq:gjsis_eqs:1} are the maximal integral submanifolds of the Cartan distribution restricted to~$\mathcal{E}$. This shows that $\mathcal{E}$ endowed with the Cartan distribution can be taken as the geometric object corresponding to system~\eqref{eq:gjsis_eqs:1}, we call such a manifold~$\mathcal{E}$ an \emph{equation}. An equation is generally of infinite dimension. Without loss of generality we assume that~\eqref{eq:gjsis_eqs:1} does not contain equations of zero order, in geometric language this means that the projection $\eval{\pi_{\infty,0}}_{\mathcal{E}}\colon\mathcal{E}\toJ^0(\pi)$ is surjective. It is obvious that at every point $\theta\in\mathcal{E}$ the Cartan plane is tangent to the equation, $\mathcal{C}_{\theta}\subset T_{\theta}(\mathcal{E})$, so that the dimension of the Cartan distribution on an equation is equal to~$n$, the same as on the jet space. System~\eqref{eq:gjsis_eqs:1} is a coordinate description of an equation~$\mathcal{E}$. Every equation has many different coordinate descriptions. \begin{example} \label{exmp:gjsis_eqs:1} Take the KdV equation (cf.~Example~\ref{exmp:gjsis_jets:1} from Section~\ref{sec:gjsis_jets:jet-spaces}) \begin{equation} \label{eq:gjsis_eqs:6} u_t-6uu_x-u_{xxx}=0. \end{equation} The bundle~$\pi$ here is the projection $\pi\colon\mathbb{R}^3\to\mathbb{R}^2$, $(x,t,u)\mapsto(x,t)$. The jet space $J^{\infty}(\pi)$ has coordinates $x$,~$t$,~$u$,~$u_x$,~$u_t$,~\dots\,,~$u_I$,~\dots{} The equation $\mathcal{E}\subsetJ^{\infty}(\pi)$ is given by the infinite series of equations \begin{align*} u_t&=6uu_x+u_{xxx}, \\ u_{tx}&=6u_x^2+6uu_{xx}+u_{xxxx}, \\ &\cdots \\ u_{tI}&=D_I(6uu_x+u_{xxx}), \\ &\cdots \end{align*} The Cartan distribution on~$\mathcal{E}$ is two-dimensional and generated by \begin{align*} D_x&=\frac{\partial}{\partial x} +\sum_s u_{s+1}\frac{\partial}{\partial u_s}, \\ D_t&=\frac{\partial}{\partial t} +\sum_s D_x^s(6uu_x+u_{xxx})\frac{\partial}{\partial u_s}, \end{align*} where $u_s=u_{x\dots x}$ ($s$ times). The functions $x$,~$t$,~$u_s$ can be taken to be coordinates on~$\mathcal{E}$. The system \begin{equation} \label{eq:gjsis_eqs:5} u_x-v=0, \quad v_x-w=0, \quad w_x-u_t+6uv=0 \end{equation} gives rise to the same equation $\mathcal{E}\subsetJ^{\infty}(\pi')$ with $\pi'\colon\mathbb{R}^5\to\mathbb{R}^2$, $(x,t,u,v,w)\mapsto(x,t)$. To prove that~\eqref{eq:gjsis_eqs:6} and~\eqref{eq:gjsis_eqs:5} define the same equation consider the map \begin{align*} &a\colonJ^2(\pi)\toJ^0(\pi'),\qquad &&a(x,t,u,u_x,u_t,u_{xx},u_{xt},u_{tt})=(x,t,u,u_x,u_{xx}), \\ &b\colonJ^0(\pi')\toJ^0(\pi),\qquad &&b(x,t,u,v,w)=(x,t,u). \end{align*} Let $a^{\infty}\colonJ^{\infty}(\pi)\toJ^{\infty}(\pi')$ and $b^{\infty}\colonJ^{\infty}(\pi')\toJ^{\infty}(\pi)$ be the lifts of these maps (cf.~Remark~\ref{sec:gjsis_jets:Lie-Back}). Then it is easy to see that $\eval{a^{\infty}}_{\mathcal{E}}\circ \eval{b^{\infty}}_{\mathcal{E}}=\eval{b^{\infty}}_{\mathcal{E}}\circ \eval{a^{\infty}}_{\mathcal{E}}=\eval{\mathrm{id}}_{\mathcal{E}}$. Lifts preserve the Cartan distributions, hence the maps $\eval{a^{\infty}}_{\mathcal{E}}$ and $\eval{b^{\infty}}_{\mathcal{E}}$ are isomorphisms of equations determined by~\eqref{eq:gjsis_eqs:6} and \eqref{eq:gjsis_eqs:5}. Thus, two different coordinate expressions~\eqref{eq:gjsis_eqs:6} and \eqref{eq:gjsis_eqs:5} determine the same equation~$\mathcal{E}$ included into two different jet spaces: \begin{equation*} \xymatrixcolsep{1pc} \xymatrixrowsep{0.5pc} \xymatrix{ &J^{\infty}(\pi) \\ \mathcal{E}\ar[ur]\ar[dr] \\ &J^{\infty}(\pi'). } \end{equation*} \end{example} As usual, the choice of functions~\eqref{eq:gjsis_eqs:1} should be restricted by some \emph{regularity} assumptions. Namely, we require that there exists a subset $\Sigma\subset\{D_I(F_s)\}$ of functions on~$J^{\infty}(\pi)$ such that \begin{enumerate} \item $F_s\in\Sigma$ for all $s=1$,~\dots,~$l$; \item the functions that belong to~$\Sigma$ define the equation~$\mathcal{E}$; \item the differentials $\mathinner{\!}\mathrm{d} f$ are locally linearly independent on~$\mathcal{E}$ for all~$f\in\Sigma$. \end{enumerate} We always require these conditions to be satisfied. In this review, we shall assume that an equation at hand~$\mathcal{E}\subsetJ^{\infty}(\pi)$ is globally defined by a relation~$F=0$, where $F$ is a section of an appropriate $l$-dimensional vector bundle~$\xi$ over the jet space~$J^{\infty}(\pi)$. This is always possible. Equations~\eqref{eq:gjsis_eqs:1} are local coordinate expressions for~$F=0$. Denote by~$P$ the $\mathcal{F}(\pi)$-module of sections of~$\xi$, so that $F\in P$. The above regularity conditions imply the following very useful fact: a function $f\inJ^{\infty}(\pi)$ vanishes on~$\mathcal{E}$, $\eval{f}_{\mathcal{E}}=0$, if and only if $f=\Delta(F)$ for some $\mathcal{C}$-differential operator~$\Delta\colon P\to\mathcal{F}(\pi)$. Let $\mathcal{E}=\{F=0\}$ be the equation defined by a section $F\in P$. The section $F$ is called \emph{normal} if for any $\mathcal{C}$-differential operator~$\Delta\colon P\to\mathcal{F}(\pi)$ such that $\Delta(F)=0$ we have $\eval{\Delta}_{\mathcal{E}}=0$. The equation~$\mathcal{E}\subsetJ^{\infty}(\pi)$ is called \emph{normal} if it can be defined by a normal section. \begin{example} A simple example of an abnormal equation is the system \begin{equation*} u_y-v_x=0,\qquad u_z-w_x=0,\qquad v_z-w_y=0. \end{equation*} Gauge equations, including Maxwell, Yang-Mills, and Einstein equations, are not normal as well. Such equations are beyond the scope of our review. On the other hand, the majority of equations of mathematical physics, in particular all evolution equations, are normal. \end{example} The Cartan vector fields on an equation~$\mathcal{E}$ form a Lie algebra~$\mathcal{C}\mathcal{X}(\mathcal{E})$. In the same manner as for jet spaces we define the Lie algebra of \emph{symmetries} of~$\mathcal{E}$ and spaces of \emph{the Cartan} and \emph{horizontal forms}: \begin{align*} \sym\mathcal{E}&=\X_{\scriptscriptstyle{\mathcal{C}}}(\mathcal{E})/\mathcal{C}\mathcal{X}(\mathcal{E}), \\ &\text{where }\X_{\scriptscriptstyle{\mathcal{C}}}(\mathcal{E}) =\sd{X\in\mathcal{X}(E)}{[X,\mathcal{C}\mathcal{X}(\mathcal{E})]\subset\mathcal{C}\mathcal{X}(\mathcal{E})}, \\ \La_{\scriptscriptstyle{\mathcal{C}}}^p(\mathcal{E}) &=\sd{\omega\in\Lambda^p(\mathcal{E})}{i_X(\omega) =0\quad\forall X\in\mathcal{C}\mathcal{X}(\mathcal{E})}, \\ \La_h^1(\mathcal{E}) &=\Lambda^1(\mathcal{E})/\La_{\scriptscriptstyle{\mathcal{C}}}^1(\mathcal{E}), \\ \La_h^q(\mathcal{E})&=\La_h^1(\mathcal{E})\wedge\dots\wedge\La_h^1(\mathcal{E}). \end{align*} We shall discuss them in more detail below. A morphism of equations $f\colon\mathcal{E}\to\mathcal{E}'$ is a smooth map that respects the Cartan distribution, i.e., for all points $\theta\in\mathcal{E}$ we have $f_*(\mathcal{C}_{\theta})\subset\mathcal{C}_{f(\theta)}$, where $\mathcal{C}_{\theta}$ is the Cartan plane at a point $\theta\in\mathcal{E}$. If a morphism $f\colon\mathcal{E}\to\mathcal{E}'$ is a fibre bundle and the map $f_*\colon\mathcal{C}_{\theta}\to\mathcal{C}_{f(\theta)}$ is an isomorphism of vector spaces for all points $\theta\in\mathcal{E}$ then $f$ is called a \emph{covering}. We shall discuss the theory of covering in Section~\ref{sec:gjsis_nonloc:nonlocal-theory}. \begin{example} \label{exmp:gjsis_eqs:3} For an equation~$\mathcal{E}$ we construct the quotient bundle \begin{equation*} \tau\colon\T(\mathcal{E})=T(\mathcal{E})/\mathcal{C}\to\mathcal{E}, \end{equation*} where $T(\mathcal{E})\to\mathcal{E}$ is the tangent bundle to~$\mathcal{E}$, $\mathcal{C}\subset T(\mathcal{E})$ is the Cartan distribution thought of as a subbundle of~$T(\mathcal{E})$. Given an inclusion $\mathcal{E}\subsetJ^{\infty}(\pi)$, $\pi\colon E\to M$, as above, the fibre bundle $\tau\colon\T(\mathcal{E})\to\mathcal{E}$ can be identified with the vertical bundle with respect to the projection $\mathcal{E}\to M$. Every Cartan vector field $X\in\mathcal{C}\mathcal{X}(\mathcal{E})$ can be lifted to a vector field $\tilde{X}\in\mathcal{X}(\T(\mathcal{E}))$ as follows. It suffices to define an action of~$\tilde{X}$ on fibre-wise linear functions on~$\T(\mathcal{E})$ that can be naturally identified with Cartan $1$-forms $\omega\in\La_{\scriptscriptstyle{\mathcal{C}}}^1(\mathcal{E})$. We put \begin{equation*} \tilde{X}(\omega)=L_X(\omega), \end{equation*} where $L_X$ denotes the Lie derivative. In coordinates, we have $\tilde{D_i}=D_i$, where $D_i$ in the right-hand side are the total derivatives on~$\T(\mathcal{E})$. Let us describe an inclusion of~$\T(\mathcal{E})$ to a jet space. Assume that~$\mathcal{E}\subsetJ^{\infty}(\pi)$ is given by~$F=0$. Let $\chi\colonJ^{\infty}(\pi\times_M\pi)\toJ^{\infty}(\pi)$ be defined by the projection to the first factor. Then $\T(\mathcal{E})\subsetJ^{\infty}(\pi\times_M\pi)$ is defined by equations \begin{equation*} \tilde{F}=0,\qquad\tilde{\ell}_F(\bi{v})=0, \end{equation*} where the tilde over~$F$ denotes the pullback by~$\chi$ (with the tilde for the Cartan vector fields defined above), $\bi{v}\in\tilde{\kappa}(\pi)=\mathcal{F}(\pi\times_M\pi,\pi)$ is the projection to the second factor $\Gamma(\pi\times_M\pi)\to\Gamma(\pi)$. So, $\T(\mathcal{E})$ is an equation and the vector fields of the form~$\tilde{X}$ generate the Cartan distribution on it. Thus, the bundle $\tau\colon\T(\mathcal{E})\to\mathcal{E}$ is a covering called the \emph{tangent covering} to~$\mathcal{E}$. In coordinates, we have $\bi{v}=(v^1,\dots,v^m)$ if the coordinates on $J^{\infty}(\pi\times_M\pi)$ are $x^i$,~$u^j_I$,~$v^j_I$, with $u^j_I$ and $v^j_I$ corresponding to the first and second factors, respectively. Thus, a coordinate description of~$\T(\mathcal{E})$ has the form \begin{equation*} F_s(x^i,u^j_I)=0,\qquad \sum_{\alpha,I}\frac{\partial F_j}{\partial u_I^{\alpha}}v_I^{\alpha}=0. \end{equation*} Note that if $\mathcal{E}$ is a normal equation then~$\T(\mathcal{E})$ is normal as well. \end{example} The reader will find more examples and a detailed discussion of coverings in Section~\ref{sec:gjsis_nonloc:nonlocal-theory} below. \subsection{Linearization} \label{sec:gjsis_eqs:line-diff-eqaut} Let $\mathcal{E}\subsetJ^{\infty}(\pi)$ be an equation defined by a section $F\in P$. Denote by~$\kappa$ the restriction of the module $\kappa(\pi)=\mathcal{F}(\pi,\pi)$ to~$\mathcal{E}$. The \emph{linearization} of the equation~$\mathcal{E}$ is the restriction to~$\mathcal{E}$ of the linearization of~$F$: \begin{equation*} \bar{\ell}_F=\eval{\ell_F}_{\mathcal{E}}\colon\kappa\to P. \end{equation*} We denote by bar the restriction of a $\mathcal{C}$-differential operator to~$\mathcal{E}$ and preserve the notation of modules for their restrictions. \begin{remark} The operator~$\ell_F$ is well-defined globally only if the module~$P$ has the form $P=\mathcal{F}(\pi,\pi')$. For an arbitrary module~$P$ the operator $\ell_F$ is defined only locally. But its restriction $\bar{\ell}_F$ is well-defined globally on the whole~$\mathcal{E}$. \end{remark} \begin{remark} \label{rem:gjsis_eqs:1} A $\mathcal{C}$-differential operator~$\Delta$ on~$\mathcal{E}$ is called \emph{normal} if for any $\mathcal{C}$-differential operator~$\square$ the condition $\square\circ\Delta=0$ implies $\square=0$. The linearization operator~$\bar{\ell}_F$ of an equation~$\mathcal{E}$ is normal if and only if the section~$F$ is normal. If an equation~$\mathcal{E}$ admits a normal representation, i.e., an embedding~$\mathcal{E}=\{F=0\}\subsetJ^\infty(\pi)$ such that the corresponding operator~$\bar{\ell}_F$ is normal then there always exists another representation~$\mathcal{E}=\{F'=0\}\subsetJ^\infty(\pi')$ for which~$\mathcal{E}$ acquires an evolutionary form. \end{remark} Recall that in coordinates the linearization~$\bar{\ell}_F$ has the form (cf.~\eqref{eq:gjsis_jets:20}): \begin{equation*} \bar{\ell}_F= \left\Vert\sum_I\frac{\partial F_j}{\partial u_I^\alpha}D_I \right\Vert_{\alpha=1,\dots,m.}^{j=1,\dots,l} \end{equation*} If an equation~$\mathcal{E}$ is defined by two different sections $F_1\in P_1$ an $F_2\in P_2$ in different, generally speaking, jet spaces \begin{equation*} \xymatrixcolsep{1pc} \xymatrixrowsep{0.5pc} \xymatrix{ &J^{\infty}(\pi_1) \\ \mathcal{E}\ar[ur]\ar[dr] \\ &J^{\infty}(\pi_2) } \end{equation*} then the corresponding linearizations $\bar{\ell}_{F_1}\colon\kappa_1\to P_1$ and $\bar{\ell}_{F_2}\colon\kappa_2\to P_2$ are \emph{equivalent} \cite{DudnikovSamborski:LOvSPDE,IgoninKerstenKrasilshchikVerbovetskyVitolo:VBGPDE} in the sense that there exist $\mathcal{C}$-differential operators $\alpha$, $\beta$, $\alpha'$, $\beta'$, $s_1$, and $s_2$ on~$\mathcal{E}$ \begin{equation} \label{eq:gjsis_eqs:3} \xymatrixcolsep{5pc} \xymatrix{ \kappa_1\ar[r]_{\bar{\ell}_{F_1}}\ar@<.5ex>[d]^{\alpha} &P_1\ar@<.5ex>[d]^{\alpha'}\ar@/_1pc/@<-1ex>[l]_{s_1} \\ \kappa_2\ar@<.5ex>[u]^{\beta}\ar[r]^{\bar{\ell}_{F_2}} &P_2\ar@<.5ex>[u]^{\beta'}\ar@/^1pc/@<1ex>[l]^{s_2} } \end{equation} such that \begin{equation*} \bar{\ell}_{F_1}\beta=\beta'\bar{\ell}_{F_2},\quad \bar{\ell}_{F_2}\alpha=\alpha'\bar{\ell}_{F_1},\quad \beta\alpha=\mathrm{id}+s_1\bar{\ell}_{F_1},\quad \alpha\beta=\mathrm{id}+s_2\bar{\ell}_{F_2}. \end{equation*} \begin{example} Consider two presentations~\eqref{eq:gjsis_eqs:6} and~\eqref{eq:gjsis_eqs:5} of the KdV equation from Example~\ref{exmp:gjsis_eqs:1}. The operators of diagram~\eqref{eq:gjsis_eqs:3} are: \begin{align*} \bar{\ell}_{F_1}&=D_t-D_x^3-6uD_x-6u_x, \\ \bar{\ell}_{F_2}&=\begin{pmatrix} D_x & -1 & 0 \\ 0 & D_x & -1 \\ -D_t+6v & 6u & D_x \end{pmatrix} \\ \intertext{and} \alpha&=\begin{pmatrix}1 \\ D_x \\ D_{xx}\end{pmatrix}, \\ \beta&=\begin{pmatrix}1 & 0 & 0\end{pmatrix}, \\ \alpha'&=\begin{pmatrix}\hphantom{-}0 \cr \hphantom{-}0 \cr -1 \end{pmatrix}, \\ \beta'&=\begin{pmatrix}-D_{xx}-6u & -D_x & -1\end{pmatrix}, \\ s_1&=0, \\ s_2&=\begin{pmatrix} 0 & 0 & 0 \\ 1 & 0 & 0 \\ D_x & 1 & 0 \end{pmatrix}. \end{align*} The form of operators $\alpha$ and $\beta$ is obvious from the form of operators $a$ and $b$ in Example~\ref{exmp:gjsis_eqs:1}. The operators $\alpha'$ and $\beta'$ show how equations~\eqref{eq:gjsis_eqs:6} and \eqref{eq:gjsis_eqs:5} are obtained one from the other. \end{example} \begin{example} Let the number of dependent variables $m$ be equal to~$1$. Consider the lift $L\colonJ^{\infty}(\pi)\toJ^{\infty}(\pi)$ of the Legendre transformation \begin{equation*} \textstyle \eval{L}_{J^1(\pi)}(x^i,u,u_{x^i})=(u_{x^i}, \sum_{\alpha}x^{\alpha}u_{x^{\alpha}}-u,x^i). \end{equation*} It is defined wherever $\det\left\Vert u_{x^ix^j}\right\Vert\ne0$ and preserves the Cartan distribution. Of course, $L$ is not a map of fibre bundles. Consider an equation~$\mathcal{E}$ defined by $F=0$, then $L^*(F)=0$ defines the same equation~$\mathcal{E}$. The operators in diagram~\eqref{eq:gjsis_eqs:3} are as follows: \begin{align*} F_1&=F, \\ F_2&=L^*(F) \\ \intertext{and} \alpha&=-1,\qquad \alpha'=1,\qquad s_1=0, \\ \beta&=-1,\qquad \beta'=1,\qquad s_2=0. \end{align*} Let us explain how to compute the maps~$\alpha$ and~$\beta$. To this end, we have to take a symmetry of the Cartan distribution $X\in\X_{\scriptscriptstyle{\mathcal{C}}}(\pi)/\mathcal{C}\mathcal{X}(\pi)$ and see how the generating section of~$X$ transforms under the Legendre map. The generating section can be computed by the formula $\phi=\omega(X)$, where $\omega=\mathinner{\!}\mathrm{d} u-\sum_i u_{x^i}\mathinner{\!}\mathrm{d} x^i$ is a Cartan form. So, $\alpha(\phi)=\omega(L(X))=L^*(\omega)(X)=-\omega(X)=-\phi$. The same holds for~$\beta$. \end{example} \subsection{Symmetries and recursions} We have defined symmetries of a differential equation~$\mathcal{E}$ as elements of quotient space \begin{equation*} \sym\mathcal{E}=\X_{\scriptscriptstyle{\mathcal{C}}}(\mathcal{E})/\mathcal{C}\mathcal{X}(\mathcal{E}), \end{equation*} where $\X_{\scriptscriptstyle{\mathcal{C}}}(\mathcal{E}) =\sd{X\in\mathcal{X}(\mathcal{E})}{[X,\mathcal{C}\mathcal{X}(\mathcal{E})] \subset\mathcal{C}\mathcal{X}(\mathcal{E})}$, that is, symmetries of equations are symmetries of the Cartan distribution on it modulo trivial symmetries (the ones belonging to the Cartan distribution). Obviously, symmetries of an equation form a Lie algebra with respect to the commutator. Given an inclusion $\mathcal{E}\subsetJ^{\infty}(\pi)$, each symmetry~$X\in\sym\mathcal{E}$ contains exactly one vector filed~$X^v\in\X_{\scriptscriptstyle{\mathcal{C}}}(\mathcal{E})$ vertical with respect to the projection $\pi_{\infty}\colon\mathcal{E}\to M$, where $M$ is the base manifold of the bundle~$\pi$. Next, for every such a field~$X^v$ there always exists an evolutionary vector field~$\bi{E}_{\phi'}$ such that \begin{equation*} X^v=\eval{\bi{E}_{\phi'}}_{\mathcal{E}}. \end{equation*} For any such a field, the restriction~$\phi=\eval{\phi'}_{\mathcal{E}}\in\kappa=\kappa\eval{\kappa(\pi)}_{\mathcal{E}}$ depends on~$X^v$ only. Hence, we can denote the field $X^v$ by~$\bi{E}_\phi$. The element~$\phi\in\kappa$ such that~$X^v=\bi{E}_{\phi}$ is called the \emph{generating section} (or a \emph{characteristic}) of this symmetry. If the equation at hand~$\mathcal{E}$ is defined by an equality $F=0$, then the existence of a symmetry~$\bi{E}_\phi$ boils down to the condition \begin{equation*} \eval{\bi{E}_{\phi'}(F)}_{\mathcal{E}}=0, \end{equation*} where, as above, $\phi'\in\kappa(\pi)$ is an arbitrary extension of~$\phi$, which is equivalent to the condition \begin{equation} \label{eq:gjsis_eqs:4} \bar{\ell}_F(\phi)=0\quad\text{on~$\mathcal{E}$.} \end{equation} This is the determining equation for the symmetries of the equation~$\mathcal{E}=\{F=0\}$. The Jacobi bracket~\eqref{eq:gjsis_jets:15} yields a Lie algebra structure on~$\sym\mathcal{E}$ in terms of generating sections: \begin{equation} \label{eq:gjsis_eqs:8} \{\phi,\psi\}=\bar{\ell}_\psi(\phi)-\bar{\ell}_\phi(\psi),\qquad \phi,\psi\in\ker\bar{\ell}_F\subset\kappa. \end{equation} Searching symmetries, that is solving~\eqref{eq:gjsis_eqs:4}, one usually begins with choosing \emph{internal coordinates} on~$\mathcal{E}$. \begin{example} We have seen in Example~\ref{exmp:gjsis_eqs:1} that on the KdV equation \begin{equation*} u_t-6uu_x-u_{xxx}=0 \end{equation*} the functions $x$,~$t$,~$u_s=u_{x\dots x}$ ($s$ times) can be taken to be coordinates on~$\mathcal{E}$. These are internal coordinates for the KdV equation. \end{example} Of course, the choice of internal coordinates is not unique. Next, one can start with finding symmetries whose generating sections depend on derivatives of order less than some number~$k$, $\phi=\phi(x,t,u,u_1,\dots,u_k)$. \begin{example} \label{exmp:gjsis_eqs:2} Let us find symmetries of the KdV equation such that $\phi=\phi(x,t,u,u_1)$. The determining equation for symmetries has the form: \begin{equation} \label{eq:gjsis_eqs:7} (D_t-D_x^3-6uD_x-6u_x)\phi(x,t,u,u_1)=0. \end{equation} The left-hand side of this equation is a polynomial in $u_3$ and~$u_2$. The coefficient of the product $u_2u_3$ is equal to~$-3\phi_{u_1u_1}$, so $\phi_{u_1u_1}=0$, and we have \begin{equation*} \phi=\phi^0(x,t,u)+\phi^1(x,t,u)u_1. \end{equation*} Now, the left-hand side of~\eqref{eq:gjsis_eqs:7} is a polynomial in $u_3$,~$u_2$ and~$u_1$, with coefficient of~$u_3$ equal to $-3D_x(\phi^1)$. Hence, \begin{equation*} \phi=\phi^0(x,t,u)+\phi^1(t)u_1. \end{equation*} With such a~$\phi$, the left-hand side of~\eqref{eq:gjsis_eqs:7} does not depend on~$u_3$ any more. The coefficient of the product~$u_1u_2$ is $-3\phi^0_{uu}$, hence \begin{equation*} \phi=\phi^{00}(x,t)+u\phi^{01}(t,x)+\phi^1(t)u_1. \end{equation*} The coefficient of~$u_2$ is $-3\phi^{01}_x$, so that \begin{equation*} \phi=\phi^{00}(x,t)+u\phi^{01}(t)+\phi^1(t)u_1 \end{equation*} and the left-hand side of~\eqref{eq:gjsis_eqs:7} is a polynomial in $u_1$ and~$u$. The coefficient of the product~$uu_1$ shows that $\phi^{01}=0$. The remaining coefficients of~$u_1$ and~$u$, and the free term reveal that \begin{equation*} \phi^{00}=c_0,\qquad \phi^1=6\phi^{00}t+c_1, \end{equation*} where $c_0$ and $c_1$ are arbitrary constants. Therefore we have found two independent symmetries of the KdV equation \begin{equation} \label{eq:gjsis_eqs:9} \phi_1=u_x\quad\text{and}\quad\phi_2=6tu_x+1. \end{equation} \end{example} Let $X\in\sym\mathcal{E}=\X_{\scriptscriptstyle{\mathcal{C}}}(\mathcal{E})/\mathcal{C}\mathcal{X}(\mathcal{E})$ be a symmetry. Vector fields $Y\in\X_{\scriptscriptstyle{\mathcal{C}}}(\mathcal{E})$ belonging the equivalence class~$X$ we shall call \emph{representatives} of~$X$. As we explained above, any symmetry has a unique representative of the form~$\bi{E}_\phi$. But this representative can be not the simplest one. \begin{example} Symmetries~\eqref{eq:gjsis_eqs:9} of the KdV equation can be represented, respectively, by the fields $Y_1=-\partial/\partial x$ and $Y_2=\partial/\partial u-6t\partial/\partial x$, because \begin{equation*} \bi{E}_{u_x}-Y_1=D_x,\qquad \bi{E}_{6tu_x+1}-Y_2=6tD_x. \end{equation*} The fields $Y_1$ and $Y_2$ are the lifts from the zero order jet space, hence they correspond to one-parameter groups of transformations: \begin{align*} Y_1\colon & x'=x-\epsilon &\text{(translation along~$x$)}, \\ Y_2\colon & x'=x-6\epsilon t,\quad u'=u+\epsilon &\text{(Galilean symmetry)}. \end{align*} \end{example} A symmetry~$X$ is called \emph{classical} if it can be represented by a field~$\eval{Y}_{\mathcal{E}}$, with $Y\in\mathcal{X}(\pi)$ being the lift from a finite order jet space. Classical symmetries form a subalgebra of the Lie algebra $\sym\mathcal{E}$. By the Lie-B\"{a}cklund theorem (see~~\cite{KrasilshchikVinogradov:SCLDEqMP}), $Y$ is a lift from the zero order jet space (if the number of dependent variables $m>1$) or from the first order jet space (if the number of dependent variables $m=1$). Thus, in coordinate language, the generating section of a classical symmetry has the form \begin{equation*} \label{eq:gjsis_eqs:10} \phi=\begin{cases}(\phi_1,\dots,\phi_m) &\text{for $m>1$}, \\ \phi(x^i,u,u_i) &\text{for $m=1$}, \end{cases} \end{equation*} where $\phi_\alpha=b_{\alpha}(x^i,u^j)+\sum_{k=1}^na_k(x^i,u^j)u^{\alpha}_k$, while $\phi$ is an arbitrary smooth function. The vector field $Y$ on~$J^0(\pi)$ or~$J^1(\pi)$ that represents the symmetry with generating section~$\phi$ has the form \begin{equation*} Y=\begin{cases}\textstyle\sum_{j=1}^mb_j\frac{\partial}{\partial u^j} -\sum_{k=1}^na_k\frac{\partial}{\partial x^k} &\text{for $m>1$}, \\ -\textstyle\sum_{i=1}^n\frac{\partial\phi}{\partial u_i}\frac{\partial}{\partial x^i}+ (\phi-\sum_{i=1}^nu_i\frac{\partial\phi}{\partial u_i}) \frac{\partial}{\partial u} +\sum_{i=1}^n(\frac{\partial\phi}{\partial x^i}+ u_i\frac{\partial\phi}{\partial u})\frac{\partial}{\partial u_i} &\text{for $m=1$}. \end{cases} \end{equation*} \begin{example} In Example~\ref{exmp:gjsis_eqs:2} above we computed symmetries of the KdV equation with generating sections depending on $x$,~$t$,~$u$, and~$u_x$. To find all classical symmetries, we allow also dependence on~$u_t$ and compute symmetries in the same manner as in Example~\ref{exmp:gjsis_eqs:2}. We get two additional classical symmetries: \begin{equation*} \phi_3=u_t\quad\text{and}\quad\phi_4=xu_x+3tu_t+2u, \end{equation*} with the corresponding one-parameter groups of transformations being \begin{align*} \phi_3\colon & t'=t-\epsilon &&\text{(translation along~$t$)}, \\ \phi_4\colon & x'=\mathrm{e}^{-\epsilon}x,\quad t'=\mathrm{e}^{-3\epsilon}t, \quad u'=\mathrm{e}^{2\epsilon}u &&\text{(scale symmetry).} \end{align*} \end{example} Solving~\eqref{eq:gjsis_eqs:4} for sections~$\phi$ that depend on at most $k$th order derivatives will not give us a complete description of all symmetries. We can find all classical symmetries or a slightly larger subspace of symmetries, which can be considered to be a \emph{lower estimate} of the full symmetry algebra. Letting the maximal order~$k$ be arbitrary and by solving~\eqref{eq:gjsis_eqs:4} describe the dependence of~$\phi$ on derivatives of order $k$,~$k-1$, etc. This will be an \emph{upper estimate} of the symmetry algebra. Sometimes these estimates allow to find more symmetries and, in some cases, obtain a complete description of the symmetry algebra. We refer the reader to \cite{KrasilshchikVinogradov:SCLDEqMP} for examples of such calculations. A different approach is to look for a \emph{recursion operator} that is a $\mathcal{C}$-differential operator $R\colon\kappa\to\kappa$ such that there exists another $\mathcal{C}$-differential operator~$R'$ satisfying the condition \begin{equation} \label{eq:gjsis_eqs:11} \bar{\ell}_FR=R'\bar{\ell}_F. \end{equation} Operators of the form $R=\square\bar{\ell}_F$, with $\square$ being arbitrary, enjoy~\eqref{eq:gjsis_eqs:11} for all equations, so we consider recursion operators modulo such trivial ones. Obviously, $R(\sym\mathcal{E})\subset\sym\mathcal{E}$, so that having a recursion operator we can produce an infinite number of symmetries from a given one. \begin{example}\label{sec:gjsis_eqs:ex_Heat-eq} The heat equation $u_t=u_{xx}$ has two recursion operators of the first order \begin{equation*} R_1=D_x\quad\text{and}\quad R_2=2tD_x+x \end{equation*} (and, of course, the identity operator, which is of no interest). \end{example} For nonlinear equations we often are able to find a nontrivial recursion operator only if we allow it to contain the ``integration'' operator~$D_x^{-1}$. In Section~\ref{sec:gjsis_nonloc:zero-curv-repr} we explain how to define such recursion operators in a rigorous way. \begin{example} As it was already mentioned, the KdV equation has the Lenard recursion operator \begin{equation} \label{eq:gjsis_eqs:12} R=D_x^2+4u+2u_xD_x^{-1}. \end{equation} It is obvious that $R(u_x)=u_t$ and $R^2(u_x)=R(u_t)=u_{xxxxx}+10uu_{xxx}+20u_xu_{xx}+30u^2u_x$. Thus, we have a new \emph{higher} (that is non-classical) symmetry of KdV. We can proceed in the same way and compute $R^3(u_x)$,~$R^4(u_x)$ and so on. As it was already mentioned (see Section~\ref{sec:gjsis_jets:hamilt-form}), all $R^k(u_x)$ exist, that is the computation of $D_x^{-1}$ will be possible for all~$k$. So, we have constructed an infinite set of symmetries of the KdV equation. If we try to apply the Lenard recursion operator to the other two classical symmetries, $\phi_2$ (Galilean) and $\phi_4$ (scale), we get $R(\phi_2)=2\phi_4$, but $R(\phi_4)$ does not exist. In fact, $R(\phi_4)$ is a \emph{nonlocal} symmetry, as explained below in Section~\ref{sec:gjsis_nonloc:nonlocal-theory}. \end{example} Now, let us explain how to compute recursion operators. To this end, note that $\mathcal{C}$-differential operators $\kappa\to\kappa$ modulo operators of the form $\square\bar{\ell}_F$ can be naturally identified with elements of $\La_{\scriptscriptstyle{\mathcal{C}}}^1(\mathcal{E})\otimes_{\mathcal{F}(\mathcal{E})}\kappa$, that is with $\kappa$-valued Cartan $1$-forms. This identification takes an operator $R\colon\kappa\to\kappa$ to the form $\omega_R\in\La_{\scriptscriptstyle{\mathcal{C}}}^1(\mathcal{E})\otimes_{\mathcal{F}(\mathcal{E})}\kappa$, such that $\omega_R(\bi{E}_{\phi})=R(\phi)$. Next, recall that the Cartan $1$-forms $\omega\in\La_{\scriptscriptstyle{\mathcal{C}}}^1(\mathcal{E})$ are functions on the tangent covering $\T(\mathcal{E})$ (see Example~\ref{exmp:gjsis_eqs:3}) that are linear along the fibres of the projection $\tau\colon\T(\mathcal{E})\to\mathcal{E}$. Hence the forms $\omega\in\La_{\scriptscriptstyle{\mathcal{C}}}^1(\mathcal{E})\otimes_{\mathcal{F}(\mathcal{E})}\kappa$ are elements of the pullback~$\tilde{\kappa}$ of the module~$\kappa$ by~$\tau$. Thus, to a recursion operator~$R\colon\kappa\to\kappa$ we assign a fibre-wise linear element $\omega_R\in\tilde{\kappa}$. In coordinates, to the operator $R=\left\Vert\sum_If_I^{\alpha\beta}D_I\right\Vert$ there corresponds the element \begin{equation*} \omega_R=\Bigl(\sum_{\beta,I}f_I^{1\beta}v_I^{\beta},\dots, \sum_{\beta,I}f_I^{m\beta}v_I^{\beta}\Bigr) \end{equation*} with $v_I^\beta$ being coordinates along the fibres of~$\tau$. Condition~\eqref{eq:gjsis_eqs:11} on $R$ is equivalent to the following condition on~$\omega_R$: \begin{equation} \label{eq:gjsis_eqs:13} \eval{\tilde{\ell}_F(\omega_R)}_{\T(\mathcal{E})}=0. \end{equation} \begin{example} For the heat equation $u_t=u_{xx}$ the tangent covering is given by \begin{equation*} u_t=u_{xx},\qquad v_t=v_{xx}. \end{equation*} Let $x$, $t$, $u_k=u_{x\dots x}$, and $v_k=v_{x\dots x}$ be internal coordinates on it. To compute recursion operators of order~$1$ in~$D_x$ we are to solve the equation \begin{equation*} (D_t-D_x^2)(f(x,t,u_k)v_x+g(x,t,u_k)v)=0. \end{equation*} It is easy to show that the solutions are: $v$, $v_x$, and $2tv_x+xv$, with the corresponding operators $\mathrm{id}$, $D_x$, and $2tD_x+x$, as it was indicated in Example~\ref{sec:gjsis_eqs:ex_Heat-eq}. \end{example} \begin{example} The tangent covering over the KdV equation has the form \begin{equation*} u_t-u_{xxx}-6uu_x=0,\qquad v_t-v_{xxx}-6uv_x-6u_xv=0. \end{equation*} Computation of recursion operators amounts to solving equations like the following: \begin{equation} \label{eq:gjsis_eqs:16} (D_t-D_x^3-6uD_x-6u_x)(f_2v_{xx}+f_1v_x+f_0v+f_{-1}v_{-1}), \end{equation} where $f_i=f_i(x,t,u_k)$ and $v_{-1}$ is a new variable such that \begin{align} D_x(v_{-1})&=v\label{eq:gjsis_eqs:14}, \\ D_t(v_{-1})&=v_{xx}+6uv.\label{eq:gjsis_eqs:15} \end{align} As we saw above, the variables~$v_I$ correspond to the total derivatives~$D_I$, so that the variable~$v_{-1}$ is defined in~\eqref{eq:gjsis_eqs:14} to correspond to~$D_x^{-1}$, while~\eqref{eq:gjsis_eqs:15} provides the equality $D_t(D_x(v_{-1}))=D_x(D_t(v_{-1}))$. We shall defer a rigorous explanation of this computation until Section~\ref{sec:gjsis_nonloc:nonlocal-theory}. One can check that $v_{xx}+4uv+2u_xv_{-1}$ is a solution of~\eqref{eq:gjsis_eqs:16}, it yields the Lenard recursion operator~\eqref{eq:gjsis_eqs:12}. \end{example} The Lie algebra structure on $\ker\bar{\ell}_F\subset\kappa$ (the Jacobi bracket on symmetries) has a natural extension to a Lie superalgebra on $\ker\eval{\tilde{\ell}_F}_{\La_{\scriptscriptstyle{\mathcal{C}}}^*(\mathcal{E}) \otimes_{\mathcal{F}(\mathcal{E})}\kappa}$, called the \emph{Nijenhuis} (also \emph{Fr\"olicher-Nijenhuis\textup{)} bracket} and denoted by $[\![\cdot\,,\cdot]\!]$. For a detailed definition we refer the reader to~\cite{KrasilshchikKersten:SROpCSDE}. Since we identified recursion operators with elements of $\ker\eval{\tilde{\ell}_F}_{\La_{\scriptscriptstyle{\mathcal{C}}}^1(\mathcal{E})\otimes_{\mathcal{F}(\mathcal{E})}\kappa}$, we can compute the Nijenhuis bracket of them. Two particular cases are of importance for us here: the bracket of a recursion operator with itself (the Nijenhuis torsion): \begin{align*} \frac12[\![ R,R,]\!](\phi_1,\phi_2 &=\{R(\phi_1),R(\phi_2)\}-R\{R(\phi_1),\phi_2\}-R\{\phi_1,R(\phi_2)\} +R^2\{\phi_1,\phi_2\} \\ &=\ell_{R,\phi_2}(R(\phi_1))-\ell_{R,\phi_1}(R(\phi_2)) +R\ell_{R,\phi_1}(\phi_2)-R\ell_{R,\phi_2}(\phi_2), \end{align*} where $\phi_1$,~$\phi_2\in\kappa$, and the bracket of a recursion operator and a symmetry with generating section $\phi\in\ker\bar{\ell}_F\subset\kappa$ (the Lie derivative): \begin{equation*} L_{\phi}(R)(\phi')=[\![ \phi,R]\!](\phi') =\{\phi,R(\phi')\}-R\{\phi,\phi'\},\qquad\phi'\in\kappa, \end{equation*} or \begin{equation*} L_{\phi}(R)=\bi{E}_{\phi}(R)-[\ell_{\phi},R]. \end{equation*} A recursion operator~$R$ is called \emph{Nijenhuis} (or \emph{hereditary}) if its Nijenhuis torsion vanishes, i.e., $[\![ R,R]\!]=0$. Almost all known recursion operators are Nijenhuis, including the operators encountered above. The main property of Nijenhuis operators is the following: for every two symmetries with generating sections $\phi_1$,~$\phi_2\in\kappa$ such that $\{\phi_1,\phi_2\}=0$, $L_{\phi_1}(R)=0$, $L_{\phi_2}(R)=0$ and for arbitrary $k_1$ and $k_2$ we have $\{R^{k_1}(\phi_1),R^{k_2}(\phi_2)\}=0$. \begin{example} For the KdV equation take $\phi_1=\phi_2=u_x$. Obviously, we have $L_{u_x}(R)=0$, where $R$ is the Lenard recursion operator~\eqref{eq:gjsis_eqs:12}, so that all symmetries $R^k(u_x)$ commute. \end{example} \subsection{Conservation laws} A \emph{conserved current}~$\omega$ on an equation~$\mathcal{E}$ with $n$ independent variables is a closed horizontal $(n-1)$-form on~$\mathcal{E}$, i.e., a form~$\omega\in\La_h^{n-1}(\mathcal{E})$ such that $\rmd_h\omega=0$. \begin{example} \label{exmp:gjsis_eqs:4} Consider the equation of continuity in fluid dynamics \begin{equation*} \rho_t+(\rho v^1)_{x_1}+(\rho v^2)_{x_2}+(\rho v^3)_{x_3}=0. \end{equation*} The form $\omega=\rho\mathinner{\!}\mathrm{d} x_1\wedge\mathinner{\!}\mathrm{d} x_2\wedge\mathinner{\!}\mathrm{d} x_3 -\rho v^1\mathinner{\!}\mathrm{d} t\wedge\mathinner{\!}\mathrm{d} x_2\wedge\mathinner{\!}\mathrm{d} x_3 +\rho v^2\mathinner{\!}\mathrm{d} t\wedge\mathinner{\!}\mathrm{d} x_1\wedge\mathinner{\!}\mathrm{d} x_3 -\rho v^3\mathinner{\!}\mathrm{d} t\wedge\mathinner{\!}\mathrm{d} x_1\wedge\mathinner{\!}\mathrm{d} x_2$ is a conserved current. \end{example} \begin{example} \label{exmp:gjsis_eqs:5} For the KdV equation $u_t-u_{xxx}-6uu_x=0$ the forms \begin{align*} &u\mathinner{\!}\mathrm{d} x+(u_{xx}+3u^2)\mathinner{\!}\mathrm{d} t, \\ &u^2\mathinner{\!}\mathrm{d} x+(2uu_{xx}-u_x^2+4u^3)\mathinner{\!}\mathrm{d} t, \\ &(u_x^2/2-u^3)\mathinner{\!}\mathrm{d} x+(u_xu_{xxx}-u_{xx}^2/2-3u^2u_{xx} +6uu_x^2-9u^4/2)\mathinner{\!}\mathrm{d} t \end{align*} are conserved currents. \end{example} For any $\eta\in\La_h^{n-2}(\mathcal{E})$ the form $\omega=\rmd_h\eta$ is always a conserved current. Such currents are called \emph{trivial} because they are not related to the equation properties and hence are of no interest. Consider the quotient space \begin{equation*} H_h^{n-1}(\mathcal{E})=\sd{\omega\in\La_h^{n-1}(\mathcal{E})}{\rmd_h\omega=0} \big/\sd{\omega\in\La_h^{n-1}(\mathcal{E})}{\omega=\rmd_h\eta, \quad\eta\in\La_h^{n-2}(\mathcal{E})} \end{equation*} of \emph{horizontal cohomology} of~$\mathcal{E}$. We also want to quotient out the \emph{topological} conserved currents that lie in the image of the map $\zeta\colon H^{n-1}(\mathcal{E})\to H_h^{n-1}(\mathcal{E})$ induced by the natural projection $\Lambda^{n-1}(\mathcal{E})\to\La_h^{n-1}(\mathcal{E})$; here $H^{n-1}(\mathcal{E})$ is the $(n-1)$st group of the de~Rham cohomology of the space~$\mathcal{E}$. Such currents are related to the topology of the equation~$\mathcal{E}$ only. Thus, we define $\cl(\mathcal{E})=H_h^{n-1}(\mathcal{E})/\im\zeta$ to be the set of \emph{conservation laws} of equation~$\mathcal{E}$. We shall now discuss how to compute conservation laws for \emph{normal} equations. To this end, let us consider the following complex: \begin{equation} \label{eq:gjsis_eqs:17} 0\xrightarrow{}\Omega^0(\mathcal{E}) \xrightarrow{\delta}\Omega^1(\mathcal{E}) \xrightarrow{\delta}\Omega^2(\mathcal{E}) \xrightarrow{\delta}\dots, \end{equation} where \begin{equation} \label{eq:gjsis_eqs:20} \Omega^0(\mathcal{E})=\cl(\mathcal{E}),\qquad \Omega^p(\mathcal{E}) =Z^p\big/\rmd_h(\La_{\scriptscriptstyle{\mathcal{C}}}^p(\mathcal{E})\otimes_{\mathcal{F}(\mathcal{E})} \La_h^{n-2}(\mathcal{E})), \end{equation} and $Z^p=\ker\rmd_h\subset\La_{\scriptscriptstyle{\mathcal{C}}}^p(\mathcal{E})\otimes_{\mathcal{F}(\mathcal{E})} \La_h^{n-1}(\mathcal{E})$, for $p>0$. The differential~$\delta$ is induced by the Cartan differential~$\rmd_{\scriptscriptstyle{\mathcal{C}}}$, so that $\delta^2=0$. \begin{remark} Complex~\eqref{eq:gjsis_eqs:17} is a part of Vinogradov's $\mathcal{C}$-spectral sequence (see \cite{Vinogradov:AlGFLFT,Vinogradov:SSAsNDEqAlGFLFTC,Vinogradov:SSLFCLLTNT,Vinogradov:CAnPDEqSC}), namely $\Omega^p(\mathcal{E}) =E_1^{p,n-1}(\mathcal{E})$ for $p>0$, $\Omega^0(\mathcal{E}) =E_1^{0,n-1}(\mathcal{E})\big/H^{n-1}(\mathcal{E})$, and $\delta=d_1^{*,n-1}$. \end{remark} Assume that $\mathcal{E}$ is a normal equation defined by a normal section $F\in P$. In this case complex~\eqref{eq:gjsis_eqs:17} can be described in the following way: \begin{equation} \label{eq:gjsis_eqs:23} \Omega^p(\mathcal{E})=\Theta^p\big/\Theta^p_{\ell}, \end{equation} where $\Theta^p$ is a subset of $\CDiff^{\mathrm{sk}}_{p-1}(\kappa,\hat{P})$ that consists of operators $\Delta\in\CDiff^{\mathrm{sk}}_{p-1}(\kappa,\hat{P})$ such that \begin{equation*} \bar{\ell}_F^*\Delta(\phi_1,\dots,\phi_{p-1}) -\sum_{\alpha=1}^{p-1}\Delta^{*_{\alpha}} (\phi_1,\dots,\bar{\ell}_F(\phi_{\alpha}),\dots,\phi_{p-1})=0 \end{equation*} for all $\phi_1,\dots,\phi_{p-1}\in\kappa$, where~${^{*_{\alpha}}}$ denotes the operation of taking the adjoint with respect to the $\alpha$th argument. The subset $\Theta^p_{\ell}\subset\Theta^p$ consists of operators $\Delta\in\Theta^p$ of the form \begin{equation} \label{eq:gjsis_eqs:22} \Delta(\phi_1,\dots,\phi_{p-1})=\sum_{\alpha=1}^{p-1}(-1)^{\alpha+1}\Delta' (\bar{\ell}_F(\phi_{\alpha}),\phi_1,\dots, \hat{\phi}_{\alpha},\dots,\phi_{p-1}) \end{equation} for some $\mathcal{C}$-differential operators $\Delta'\colon P\times\kappa\times\dots\times\kappa\to\hat{P}$. In particular, for $p=1$, we have \begin{equation} \label{eq:gjsis_eqs:18} \Omega^1(\mathcal{E})=\ker\bar{\ell}_F^*\subset\hat{P}. \end{equation} If $p=2$ then \begin{equation*} \Omega^2(\mathcal{E}) =\sd{\Delta\in\CDiff(\kappa,\hat{P})} {\bar{\ell}_F^*\Delta=\Delta^*\bar{\ell}_F}\big/ \sd{\Delta'\bar{\ell}_F} {\Delta'\in\CDiff(P,\hat{P}),\ {\Delta'}^*=\Delta'}. \end{equation*} The differential~$\delta\colon\Omega^0(\mathcal{E})\to\Omega^1(\mathcal{E})$ is given by the formula $\delta(\omega)=\eval{\nabla^*(1)}_{\mathcal{E}}$, with~$\nabla$ being a $\mathcal{C}$-differential operator from~$P$ to~$\La_h^n(\pi)$ such that $\rmd_h\omega=\nabla(F)$ on~$J^{\infty}(\pi)$. The differential~$\delta\colon\Omega^p(\mathcal{E})\to\Omega^{p+1}(\mathcal{E})$, $p\geq1$, has the form \begin{equation*} \delta(\Delta)(\phi_1,\dots,\phi_p) =\sum_{\alpha=1}^p(-1)^{\alpha+1} \ell_{\Delta,\phi_1,\dots,\hat{\phi}_{\alpha},\dots,\phi_p}(\phi_{\alpha}) +{\eval{\nabla}_{\mathcal{E}}}^{*_1}(\phi_1,\dots,\phi_p), \end{equation*} where~$\nabla$ is a $\mathcal{C}$-differential operator~$\nabla\colon P\times\kappa(\pi)\times\dots\times\kappa(\pi)\to\hat{\kappa}(\pi)$ that satisfies the relation \begin{equation} \label{eq:gjsis_eqs:21} \ell_F^*\Delta(\phi_1,\dots,\phi_{p-1}) -\sum_{\alpha=1}^{p-1}\Delta^{*_{\alpha}} (\phi_1,\dots,\ell_F(\phi_{\alpha}),\dots,\phi_{p-1}) =\nabla(F,\phi_1,\dots,\phi_{p-1}) \end{equation} on~$J^{\infty}(\pi)$. In the case of an evolution equation $F=u_t-f$ we have $\Delta\in\CDiff^{\mathrm{sk\,*}}(\kappa,\hat{\kappa})$ and the operator~$\nabla$ can be chosen in the form \begin{equation*} \nabla(h,\phi_1,\dots,\phi_{p-1}) =-\ell_{\Delta,\phi_1,\dots,\phi_{p-1}}(h), \end{equation*} where $h\in P$. If $p=1$ and $\psi\in\im\delta\subset\Omega^1(\mathcal{E})$ then we can put $\nabla=-\ell_{\psi}^*$. The above description of elements in~$\Omega^p(\mathcal{E})$ in terms of $\mathcal{C}$-differential operators makes sense only for a given inclusion of the equation~$\mathcal{E}$ to a jet space. If we consider two inclusions \begin{equation*} \xymatrixcolsep{1pc} \xymatrixrowsep{0.5pc} \xymatrix{ &J^{\infty}(\pi_1) \\ \mathcal{E}\ar[ur]\ar[dr] \\ &J^{\infty}(\pi_2), } \end{equation*} so that the corresponding linearizations are equivalent and we have diagram~\eqref{eq:gjsis_eqs:3}, then the operators $\Delta_1$ and $\Delta_2$ that define the same element of~$\Omega^p(\mathcal{E})$ with respect to two inclusions are related as follows: \begin{equation*} \Delta_1={\alpha'}^*\Delta_2 (\alpha(\,\cdot\,),\dots,\alpha(\,\cdot\,)),\qquad \Delta_2={\beta'}^*\Delta_1 (\beta(\,\cdot\,),\dots,\beta(\,\cdot\,)). \end{equation*} Elements of~$\Omega^1(\mathcal{E})$ are said to be \emph{cosymmetries} of the equation~$\mathcal{E}$. We say that isomorphism~\eqref{eq:gjsis_eqs:18} takes a cosymmetry to its \emph{generating section} (or the \emph{characteristic}) belonging to~$\ker\bar{\ell}_F^*\subset\hat{P}$. The ``Two-line Theorem'' by Vinogradov (see, e.g., \cite{KrasilshchikVinogradov:SCLDEqMP}) implies that complex~\eqref{eq:gjsis_eqs:17} is exact at the term~$\Omega^0(\mathcal{E})$, hence the conservation laws of~$\mathcal{E}$ form a subset of the space of cosymmetries, $\cl(\mathcal{E})\subset\Omega^1(\mathcal{E})$. The generating section of a cosymmetry that belongs to $\delta(\cl(\mathcal{E}))$ is called the \emph{generating section of the conservation law}. As noted above, to compute the generating section of a conservation law we extend it arbitrarily to a form $\omega\in\La_h^{n-1}(\pi)$ on the jet space $J^{\infty}(\pi)$, so that $\eval{\rmd_h\omega}_{\mathcal{E}}=0$. Hence there exists a $\mathcal{C}$-differential operator $\Delta\colon P\to\La_h^n(\pi)$ such that $\rmd_h\omega=\Delta(F)$; the element $\psi={\eval{\Delta^*}_{\mathcal{E}}}(1)\in\hat{P}$ is the generating section of the conservation law under consideration. The generating section~$\psi$ of a conservation law can always be extended to the jet space~$J^\infty(\pi)$ in such a way that $\langle\psi,F\rangle=\rmd_h\omega$, with $\omega$ being a conserved current for the same conservation law. \begin{remark} The above describe procedure is one of the ways to compute the differential~$\delta\colon\Omega^0(\mathcal{E})\to\Omega^1(\mathcal{E})$ in~\eqref{eq:gjsis_eqs:17} for an arbitrary equation~$\mathcal{E}$. If~$\mathcal{E}$ is presented in an evolutionary form then computation of the generating section is simpler and more straightforward. Namely, let~$t$, $x^1,\dots,x^n$ be the independent variables and~$[\omega]\in\Omega^0(\mathcal{E})$, where \begin{equation*} \omega=X\mathinner{\!}\mathrm{d} x^1\wedge\dots\wedge x^n+\sum_{i=1}^n T_i\mathinner{\!}\mathrm{d} t\wedge\mathinner{\!}\mathrm{d} x^1\wedge\dots\wedge\hat{\mathinner{\!}\mathrm{d} x}^i\wedge\dots\wedge\mathinner{\!}\mathrm{d} x^n. \end{equation*} Then the corresponding generating section is \begin{equation*} \psi=\left(\frac{\delta X}{\delta u^1},\dots,\frac{\delta X}{\delta u^m}\right), \end{equation*} where~$\delta/\delta u^j$ is the variational derivative with respect to~$u^j$. \end{remark} \begin{example} The generating section of the conservation law from Example~\ref{exmp:gjsis_eqs:4} is equal to ~$1$. The generating sections of the conservation laws from Example~\ref{exmp:gjsis_eqs:5} are~$1$, $2u$, and~$u_{xx}+3u^2$. \end{example} The determining equation for the conservation laws (or, to be more precise, of cosymmetries) of the equation~$\mathcal{E}=\{F=0\}$ is \begin{equation} \label{eq:gjsis_eqs:19} \bar{\ell}_F^*(\psi)=0. \end{equation} This equation is dual to equation~\eqref{eq:gjsis_eqs:4} for symmetries. The computations involved in solving~\eqref{eq:gjsis_eqs:19} are very similar to those used to compute symmetries. However, unlike the case of symmetries, not all solutions of~\eqref{eq:gjsis_eqs:19} give conservation laws. To check if the generating section of a cosymmetry corresponds to a conservation law we can use the following corollary of the ``Two-line Theorem'' by Vinogradov (see \cite{KrasilshchikVinogradov:SCLDEqMP}): if the de Rham cohomology $H^n(\mathcal{E})=0$, complex~\eqref{eq:gjsis_eqs:17} is exact at the term~$\Omega^1(\mathcal{E})$. So, in this case the generating section~$\psi\in\ker\bar{\ell}_F^*$ is the generating section of a conservation law if and only if $\delta(\psi)=0$, that is, there exists a self-adjoint operator $\Delta'={\Delta'}^*\in\mathcal{C}(P,\hat{P})$ such that \begin{equation*} \ell_{\psi}+{\eval{\nabla}_{\mathcal{E}}}^*=\Delta'\bar{\ell}_F, \end{equation*} where $\nabla\in\mathcal{C}(\kappa(\pi),\hat{P})$ satisfies the equality \begin{equation*} \ell_F^*(\psi)=\nabla(F)\quad\text{on~$J^{\infty}(\pi)$.} \end{equation*} Note that the generating section~$\psi$ of a conservation law can always be extended to the jet space $J^{\infty}(\pi)$ in such a way that the horizontal $n$-form $\langle\psi,F\rangle$ will be exact: $\langle\psi,F\rangle=\rmd_h\omega$, with $\eval{\omega}_{\mathcal{E}}$ being a conserved current that corresponds to~$\psi.$ \subsection{A parallel with finite-dimensional differential geometry. III} \label{sec:gjsis_eqs:parallel-with-finite} In this section we begin compilation of a dictionary between the geometry of normal differential equations and finite-dimensional differential geometry, similar to the one for jet spaces from Sections~\ref{sec:gjsis_jets:parallel-with-finite} and~\ref{sec:gjsis_jets:parallel-with-finite-1}. We start just as we did for jet spaces: we consider the space~$\mathcal{E}$ endowed with the Cartan distribution~$\mathcal{C}$ and take for the points of that ``manifold'' the maximal integral submanifolds of~$\mathcal{C}$, i.e., the solutions of~$\mathcal{E}$. Further, the dictionary reads: \begin{eqnarray*} \text{\textbf{Manifold~$M$}}&& \text{\textbf{Normal equation~$\mathcal{E}$}} \\[1ex] \text{points}&\quad\longleftrightarrow\quad&\text{solutions} \\ \text{functions~$C^{\infty}(M)$}&\quad\longleftrightarrow\quad& \text{conservation laws~$\cl(\mathcal{E})$} \\ \text{the de Rham complex}&\quad\longleftrightarrow\quad& \text{complex~\eqref{eq:gjsis_eqs:17}}\\[-1.5ex] \text{of differential forms}&\quad\hphantom{\longleftrightarrow}\quad& \dots\xrightarrow{}\Omega^{p-1}(\mathcal{E}) \xrightarrow{\delta}\Omega^p(\mathcal{E}) \xrightarrow{}\dots \\ \text{vector fields}&\quad\longleftrightarrow\quad&\text{symmetries} \\ \text{the tangent bundle}&\quad\longleftrightarrow\quad& \text{the tangent covering~$\tau\colon\T(\mathcal{E})\to\mathcal{E}$} \end{eqnarray*} In addition to considerations from Section~\ref{sec:gjsis_jets:parallel-with-finite} on jet spaces, this dictionary is justified by the following facts. First, on a finite-dimensional manifold the differential forms are functions on the tangent bundle with odd fibres. Correspondingly, definition~\eqref{eq:gjsis_eqs:20} shows that \begin{equation*} \Omega^*(\mathcal{E})=\cl(\T(\mathcal{E})), \end{equation*} with elements of~$\Omega^p(\mathcal{E})$ given by fibre-wise $p$-linear conserved currents. Fibres of the tangent covering $\tau\colon\T(\mathcal{E})\to\mathcal{E}$ are assumed to be odd. Since elements of~$\Omega^p(\mathcal{E})$ can be understood as conservation laws on~$\T(\mathcal{E})$, we can ask what are the generating section of these conservation laws? For the element of~$\Omega^p(\mathcal{E})$ that corresponds to an operator $\Delta\in\CDiff^{\mathrm{sk}}_{p-1}(\kappa,\hat{P})$ the generating section is $(-\nabla^{*_1},\Delta)$, where the operator~$\nabla$ is given by~\eqref{eq:gjsis_eqs:21}. Here we interpret skew-symmetric $\mathcal{C}$-differential operators $\kappa\times\dots\times\kappa\to Q$, modulo operators of the form~\eqref{eq:gjsis_eqs:22}, as elements of the module~$Q$ pulled back on~$\T(\mathcal{E})$. This describes the isomorphism~\eqref{eq:gjsis_eqs:23}. Second, on a finite-dimensional manifold two natural actions of vector fields on differential forms exist: the interior product and the Lie derivative. Correspondingly, on an equation~$\mathcal{E}$ the evolution field~$\bi{E}_{\phi}$, $\phi\in\sym\mathcal{E}$, induces the interior product and the Lie derivative on~$\Omega^*(\mathcal{E})$ defined by~\eqref{eq:gjsis_eqs:20}: \begin{equation*} i_{\phi}\colon\Omega^p(\mathcal{E})\to\Omega^{p-1}(\mathcal{E}),\qquad L_{\phi}\colon\Omega^p(\mathcal{E})\to\Omega^{p+1}(\mathcal{E}). \end{equation*} These operations are related to the differential~$\delta$ by the usual identity \begin{equation*} L_{\phi}=\delta i_{\phi}+i_{\phi}\delta. \end{equation*} In terms of $\mathcal{C}$-differential operators, the interior product $i_{\phi}\colon\Omega^p(\mathcal{E})\to\Omega^{p-1}(\mathcal{E})$ for $p>1$ is the contraction of the operator with~$\phi$. For $p=1$, the interior product $i_{\phi}(\psi)$, $\psi\in\ker\bar{\ell}_F^*$, arises from the Green formula and is the conservation law defined by the conserved current~$\eval{\omega}_{\mathcal{E}}\in\La_h^{n-1}(\mathcal{E})$ such that \begin{equation*} \langle\ell_F(\phi),\psi\rangle-\langle\phi,\ell_F^*(\psi)\rangle =\rmd_h\omega \quad\text{on~$J^{\infty}(\pi)$}. \end{equation*} If $\psi\in\hat{P}$ is a generating section of a conservation law, then a symmetry~$\phi\in\sym\mathcal{E}$ acts on it by the formula \begin{equation*} L_{\phi}(\psi)=\bi{E}_{\phi}(\psi)+\square^*(\psi), \end{equation*} with some operator $\square\in\CDiff(P,P)$ that satisfies the equality~$\ell_F(\phi)=\square(F)$ on~$J^{\infty}(\pi)$. Third, on a finite-dimensional manifold a symplectic form gives rise to a Poisson bracket. The corresponding construction for an equation~$\mathcal{E}$ relies on the notion of a \emph{symplectic structure} that is a closed element of~$\Omega^2(\mathcal{E})$. We do not assume that the symplectic form is non-degenerate, so the Poisson bracket will be defined on a subset of~$\cl\mathcal{E}$ (recall that conservation laws are analogues of functions on~$\mathcal{E}$.) In terms of $\mathcal{C}$-differential operators, a symplectic structure is the equivalence class of operators $\Delta\in\CDiff(\kappa,\hat{P})$ such that \begin{equation} \label{eq:gjsis_eqs:25} \bar{\ell}_F^*\Delta=\Delta^*\bar{\ell}_F, \qquad \ell_{\Delta,\phi_2}(\phi_1)-\ell_{\Delta,\phi_1}(\phi_2) +{\eval{\nabla}_{\mathcal{E}}}^{*_1}(\phi_1,\phi_2)=0, \end{equation} where $\phi_1$,~$\phi_2\in\kappa$, $\nabla\colon P\times\kappa\to\kappa$ is a $\mathcal{C}$-differential operator such that \begin{equation*} \ell_F^*\Delta-\Delta^*\ell_F=\nabla(F,\,\cdot\,) \quad\text{on~$J^{\infty}(\pi)$}, \end{equation*} modulo operators of the form $\Delta'\bar{\ell}_F$, $\Delta'\in\CDiff(P,\hat{P})$, ${\Delta'}^*=\Delta'$. \begin{example} For the simplest WDVV equation \begin{equation*} u_{yyy}-u_{xxy}^2+u_{xxx}u_{xyy}=0 \end{equation*} the operator~$D_x$ is a symplectic structure. \end{example} For evolution equations conditions~\eqref{eq:gjsis_eqs:25} amount to \begin{equation*} \Delta^*=-\Delta, \qquad \ell_{\Delta,\phi_1}(\phi_2)-\ell_{\Delta,\phi_2}(\phi_1) =\ell_{\Delta,\phi_1}^*(\phi_2). \end{equation*} The construction of the Poisson bracket is similar to the one on a finite-dimensional manifold. Let $\Omega\in\Omega^2(\mathcal{E})$ be a symplectic structure, i.e., $\delta(\Omega)=0$. A conservation law with the generating section~$\psi$ is called \emph{admissible} if there exists a symmetry~$\phi\in\sym\mathcal{E}$ such that \begin{equation} \label{eq:gjsis_eqs:24} \psi=i_{\phi}(\Omega). \end{equation} Symmetries that correspond to admissible conservation laws in the sense of~\eqref{eq:gjsis_eqs:24} are called \emph{Hamiltonian symmetries}. By definition, \emph{the Poisson bracket of two admissible conservation laws} $\omega$ and~$\omega'$ with the generating sections $\psi$ and $\psi'$, respectively, has the generating section \begin{equation*} \{\omega,\omega'\}_{\Omega} =L_{\phi}(\psi') =i_{\phi}i_{\phi'}\Omega, \end{equation*} where $\phi$ and $\phi'$ are Hamiltonian symmetries corresponding to conservation laws $\omega$ and~$\omega'$. To the conservation law $\{\omega,\omega'\}_{\Omega}$ there corresponds the Jacobi bracket $\{\phi,\phi'\}$, so that Hamiltonian symmetries form a Lie algebra: $[L_{\phi},L_{\phi'}]=L_{\{\phi,\phi'\}}$. As we see from~\eqref{eq:gjsis_eqs:24}, a symplectic structure takes a symmetry to a conservation law, in terms of operators~\eqref{eq:gjsis_eqs:25}, this map has the form $\psi=\Delta(\phi)$. \subsection{Cotangent covering to a normal equation} \label{sec:gjsis_eqs:cotangent-covering} In the previous section we discussed geometry related to the tangent covering and functions on it (forms). But what about the cotangent covering? We have defined the tangent covering for an equation~$\mathcal{E}$ without fixing an inclusion of~$\mathcal{E}$ to a jet space. Dualizing such an invariant definition requires use of rather complicated homological algebra, so we will define the cotangent covering for an equation~$\mathcal{E}$ embedded to a jet space, $\mathcal{E}\subsetJ^{\infty}(\pi)$, and then check that the construction does not depend on the choice of the embedding. For a normal equation~$\mathcal{E}$ given by a normal section $F=0, $~$F\in P$, with $P$ being a module over~$J^{\infty}(\pi)$, we define the equation~$\Ts(\mathcal{E})\subsetJ_h^{\infty}(\hat{P})$ by the equalities \begin{equation*} \tilde{F}=0,\qquad\tilde{\ell}_F^*(\bi{p})=0, \end{equation*} where the tilde denotes the pullback to~$J_h^{\infty}(\hat{P})$ and $\bi{p}\in\tilde{\hat{P}}$ corresponds to the identity operator $\hat{P}\to\hat{P}$ under the identification $\tilde{\hat{P}}=\CDiff(\hat{P},\hat{P})$. The natural projection $\tau^*\colon\Ts(\mathcal{E})\to\mathcal{E}$ is called the \emph{cotangent covering} to~$\mathcal{E}$. In coordinates, we have $\bi{p}=(p^1,\dots,p^l)$ if the coordinates on $J_h^{\infty}(\hat{P})$ are $x^i$,~$u^j_I$,~$p^j_I$, with $u^j_I$ and $p^j_I$ being fibre coordinates along projections $\mathcal{E}\to M$ and $J_h^{\infty}(\hat{P})\to\mathcal{E}$, respectively. Below we assume that not only equation~$\mathcal{E}$ is normal but the operator $\bar{\ell}_F^*$ is normal (see Remark~\ref{rem:gjsis_eqs:1}) as well. Then $\Ts(\mathcal{E})$ will also be a normal equation. \begin{remark} Obviously, for every~$\mathcal{E}$ the cotangent equation~$\Ts(\mathcal{E})$ is an Euler-Lagrange equation with the Lagrangian density $L=\langle F,\bi{p}\rangle$. In applications, considering $\Ts(\mathcal{E})$ instead of~$\mathcal{E}$ is occasionally useful to handle the equation as though it were Lagrangian (see, e.g., \cite[Volume~1, Sections~3.2,~3.3]{MorseFeshbach:MTP}). We refer to~\cite[Section~4.5.1]{FilippovSavchinShorokhov:VPNOp} and \cite[Section~5]{PopovychKunzingerIvanova:CLPSLPEq} for more details and references. \end{remark} If we have two inclusions of equation~$\mathcal{E}$ to jet spaces \begin{equation*} \xymatrixcolsep{1pc} \xymatrixrowsep{0.5pc} \xymatrix{ &J^{\infty}(\pi_1) \\ \mathcal{E}\ar[ur]\ar[dr] \\ &J^{\infty}(\pi_2), } \end{equation*} then the adjoint linearizations $\bar{\ell}_{F_1}^*$ and $\bar{\ell}_{F_2}^*$ are equivalent: \begin{equation} \label{eq:gjsis_eqs:29} \xymatrixcolsep{5pc} \xymatrix{ \hat{P}_1\ar[r]_{\bar{\ell}_{F_1}^*}\ar@<.5ex>[d]^{{\beta'}^*} &\hat{\kappa}_1\ar@<.5ex>[d]^{\beta^*}\ar@/_1pc/@<-1ex>[l]_{s_1^*} \\ \hat{P}_2\ar@<.5ex>[u]^{{\alpha'}^*}\ar[r]^{\bar{\ell}_{F_2}^*} &\hat{\kappa}_2\ar@<.5ex>[u]^{\alpha^*}\ar@/^1pc/@<1ex>[l]^{s_2^*} } \end{equation} such that \begin{equation*} \bar{\ell}_{F_1}^*{\alpha'}^*=\alpha^*\bar{\ell}_{F_2}^*,\quad \bar{\ell}_{F_2}^*{\beta'}^*=\beta^*\bar{\ell}_{F_1}^*,\quad {\alpha'}^*{\beta'}^*=\mathrm{id}+s_1^*\bar{\ell}_{F_1}^*,\quad {\beta'}^*{\alpha'}^*=\mathrm{id}+s_2^*\bar{\ell}_{F_2}^*, \end{equation*} where the operators $\alpha$,~$\beta$,~$\alpha'$,~$\beta'$,~$s_1$, and~$s_2$ are defined in~\eqref{eq:gjsis_eqs:3}. Therefore, the cotangent coverings constructed using $\ell_{F_1}^*$ and $\ell_{F_2}^*$ are isomorphic, thus the cotangent coverings do not depend on the choice of inclusion $\mathcal{E}\subsetJ^{\infty}(\pi)$. Now we describe an isomorphism between $\sym\mathcal{E}$ and the subspace of $\cl\Ts(\mathcal{E})$ that consists of conservation laws with fibre-wise linear conserved currents. Thus we will justify the first parallel in the prolongation of our dictionary: \begin{eqnarray} \text{\textbf{Manifold~$M$}}&& \text{\textbf{Normal equation~$\mathcal{E}$}} \nonumber \\[1ex] \text{the cotangent bundle}&\quad\longleftrightarrow\quad& \text{the cotangent covering~$\tau^*\colon \Ts(\mathcal{E})\to\mathcal{E}$} \nonumber \\ \label{eq:gjsis_eqs:26} \text{multivector fields}&\quad\longleftrightarrow\quad& \text{conservation laws $\cl(\Ts(\mathcal{E}))$} \end{eqnarray} Let $\phi\in\kappa$ be the generating section of a symmetry of~$\mathcal{E}$. Extend it to an element $\phi\in\kappa(\pi)$ and consider the Green formula \begin{equation} \label{eq:gjsis_eqs:27} \langle\ell_F(\phi),\psi\rangle-\langle\phi,\ell_F^*(\psi)\rangle =\rmd_h\omega(\phi,\psi), \end{equation} where $\psi\in\hat{P}$, $\omega(\phi,\psi)\in\La_h^{n-1}(\pi)$. The mapping $\psi\mapsto\omega(\phi,\psi)$ is a $\mathcal{C}$-differential operator $\hat{P}\to\La_h^{n-1}(\pi)$, so that it gives rise to a closed form $\omega_{\phi}\in\La_h^{n-1}(\Ts(\mathcal{E}))$. The induced map $\sym\mathcal{E}\to\cl\Ts(\mathcal{E})$, which takes the symmetry with the generating functions~$\phi$ to the conservation law with the current $\omega_{\phi}$, gives the desired isomorphism. In fact, formula~\eqref{eq:gjsis_eqs:27} gives more. It holds not only for generating sections of symmetries of~$\mathcal{E}$, but also for an arbitrary $\phi\in\kappa$, so that we obtain a map $\kappa\to\La_h^{n-1}(\Ts(\mathcal{E}))$. Since $\tilde{\kappa}$ is a direct summand in $\kappa(\Ts(\mathcal{E}))$, the element $\omega(\phi,\psi)$ yields an element $\rho\in\Omega^1(\Ts(\mathcal{E}))$. One can prove that $\rho$ does not depend on the choice of inclusion $\mathcal{E}\toJ^{\infty}(\pi)$ used in its construction. In terms of isomorphism~\eqref{eq:gjsis_eqs:23}, we have $\rho=(\bi{p},0)$. This is a very important element since it plays the r\^ole of the canonical $1$-form $p\mathinner{\!}\mathrm{d} q$ on a finite-dimensional cotangent space: \begin{eqnarray*} \text{\textbf{Manifold~$M$}}&& \text{\textbf{Normal equation~$\mathcal{E}$}} \\[1ex] \text{the canonical $1$-form $p\mathinner{\!}\mathrm{d} q$}&\quad\longleftrightarrow\quad& \text{$\rho\in\Omega^1(\Ts(\mathcal{E}))$} \\ \text{the canonical symplectic form} &\quad\longleftrightarrow\quad& \text{canonical symplectic structure} \\[-1ex] \text{$\mathinner{\!}\mathrm{d} p\wedge\mathinner{\!}\mathrm{d} q$} &\quad\longleftrightarrow\quad& \text{$\Omega=\delta(\rho)\in\Omega^2(\Ts(\mathcal{E}))$} \end{eqnarray*} In terms of operators~\eqref{eq:gjsis_eqs:23}, the canonical symplectic structure on~$\Ts(\mathcal{E})$ has the form \begin{equation*} \Omega=\begin{pmatrix} 0 & 1 \\ -1 & 0. \end{pmatrix} \end{equation*} As to entry~\eqref{eq:gjsis_eqs:26} of our dictionary, we take it for the definition of multivectors. Since we are interested in skew-symmetric multivectors, we assume the fibres of the cotangent covering $\Ts(\mathcal{E})\to\mathcal{E}$ to be \emph{odd}. We call conservation laws of~$\Ts(\mathcal{E})$ whose currents are fibre-wise $p$-linear \emph{variational $p$-vectors} on~$\mathcal{E}$ and denote their set by $D_p(\mathcal{E})$. Thus, $D_0(\mathcal{E})=\cl\mathcal{E}$ and $D_1(\mathcal{E})=\sym\mathcal{E}$. For $D_p(\mathcal{E})$ we have a description in terms of $\mathcal{C}$-differential operators similar to~\eqref{eq:gjsis_eqs:23}. Namely, \begin{equation*} D_p(\mathcal{E})=\Xi_p\big/\Xi_p^{\ell}, \end{equation*} where $\Xi_p$ is a subset of $\CDiff^{\mathrm{sk}}_{p-1}(\hat{P},\kappa)$ that consists of operators $\Delta\in\CDiff^{\mathrm{sk}}_{p-1}(\hat{P},\kappa)$ such that \begin{equation*} \bar{\ell}_F\Delta(\psi_1,\dots,\psi_{p-1}) -\sum_{\alpha=1}^{p-1}\Delta^{*_{\alpha}} (\psi_1,\dots,\bar{\ell}_F^*(\psi_{\alpha}),\dots,\psi_{p-1})=0 \end{equation*} for all $\psi_1,\dots,\psi_{p-1}\in\hat{P}$, where~${^{*_{\alpha}}}$ denotes, as before, the operation of taking adjoint with respect to the $\alpha$th argument. The subset $\Xi_p^{\ell}\subset\Xi_p$ consists of operators $\Delta\in\Xi_p$ of the form \begin{equation*} \Delta(\psi_1,\dots,\psi_{p-1})=\sum_{\alpha=1}^{p-1}(-1)^{\alpha+1}\Delta' (\bar{\ell}_F^*(\psi_{\alpha}),\psi_1,\dots, \hat{\psi}_{\alpha},\dots,\psi_{p-1}) \end{equation*} for some $\mathcal{C}$-differential operators $\Delta'\colon \hat{\kappa}\times\hat{P}\times\dots\times\hat{P}\to\kappa$. In particular, for $p=2$, we have \begin{equation*} D_2(\mathcal{E}) =\sd{\Delta\in\CDiff(\hat{P},\kappa)} {\bar{\ell}_F\Delta=\Delta^*\bar{\ell}_F^*}\big/ \sd{\Delta'\bar{\ell}_F^*} {\Delta'\in\CDiff(\hat{\kappa},\kappa),\ {\Delta'}^*=\Delta'}. \end{equation*} For the element of~$D_p(\mathcal{E})$ that corresponds to an operator $\Delta\in\CDiff^{\mathrm{sk}}_{p-1}(\hat{P},\kappa)$ the generating section is $(-\nabla^{*_1},\Delta)$, where the operator~$\nabla\colon P\times\hat{P}\times\dots\times\hat{P}\to P$ is given by \begin{equation} \label{eq:gjsis_eqs:28} \ell_F\Delta(\psi_1,\dots,\psi_{p-1}) -\sum_{\alpha=1}^{p-1}\Delta^{*_{\alpha}} (\psi_1,\dots,\ell_F^*(\psi_{\alpha}),\dots,\psi_{p-1}) =\nabla(F,\psi_1,\dots,\psi_{p-1}). \end{equation} In the case of evolution equation $F=u_t-f$ we have $\Delta\in\CDiff^{\mathrm{sk\,*}}(\hat{\kappa},\kappa)$ and the operator~$\nabla$ can be chosen in the form: \begin{equation*} \nabla(h,\psi_1,\dots,\psi_{p-1}) =\ell_{\Delta,\psi_1,\dots,\psi_{p-1}}(h), \end{equation*} where $h\in P$. Since we assume the fibres of cotangent covering to be odd, the bracket defined on $\cl(\Ts(\mathcal{E}))$ by the canonical symplectic structure will be the \emph{variational Schouten bracket} \begin{equation*} [\![\,\cdot\,,\cdot\,]\!]\,\colon D_k\times D_l\to D_{k+l-1}. \end{equation*} In terms of $\mathcal{C}$-differential operators, this bracket has the form: \begin{multline*} [\![\Delta_1,\Delta_2]\!](\psi_1,\dots,\psi_{k+l-2})=\sum_{\sigma\in S_{k+l-2}^{l-1}}(-1)^\sigma \ell_{\Delta_2,\psi_{\sigma(1,l-1)}}(\Delta_1(\psi_{\sigma(l,k+l-2)})) \\ -(-1)^{(k-1)l}\sum_{\sigma\in S_{k+l-2}^k}(-1)^\sigma \Delta_2(\nabla_1^{*_1}(\psi_{\sigma(1,k)}), \psi_{\sigma(k+1,k+l-2)}) \\ -(-1)^{(k-1)(l-1)}\sum_{\sigma\in S_{k+l-2}^{k-1}}(-1)^\sigma \ell_{\Delta_1,\psi_{\sigma(1,k-1)}}(\Delta_2(\psi_{\sigma(k,k+l-2)})) \\ +(-1)^{l-1}\sum_{\sigma\in S_{k+l-2}^l}(-1)^\sigma \Delta_1(\nabla_2^{*_1}(\psi_{\sigma(1,l)}),\psi_{\sigma(l+1,k+l-2)}), \end{multline*} where $\Delta_1\in D_k(P)$, $\Delta_2\in D_l(P)$, $\nabla_1$ and $\nabla_2$ are defined by~\eqref{eq:gjsis_eqs:28}, $\psi_1$,~\dots,~$\psi_{k+l-2}\in\hat{P}$, cf.~\eqref{eq:gjsis_jets:41}. The above description of variational multivectors in terms of $\mathcal{C}$-differential operators makes sense only for a given inclusion of equation~$\mathcal{E}$ to a jet space. If we consider two inclusions \begin{equation*} \xymatrixcolsep{1pc} \xymatrixrowsep{0.5pc} \xymatrix{ &J^{\infty}(\pi_1) \\ \mathcal{E}\ar[ur]\ar[dr] \\ &J^{\infty}(\pi_2), } \end{equation*} so that the corresponding linearizations and adjoint linearizations are equivalent and we have diagrams~\eqref{eq:gjsis_eqs:3} and~\eqref{eq:gjsis_eqs:29}, then the operators $\Delta_1$ and $\Delta_2$ that define the same element of~$D_p(\mathcal{E})$ with respect to the two inclusions are related as follows: \begin{equation*} \Delta_2=\alpha\Delta_1 ({\alpha'}^*(\,\cdot\,),\dots,{\alpha'}^*(\,\cdot\,)),\qquad \Delta_1=\beta\Delta_2 ({\beta'}^*(\,\cdot\,),\dots,{\beta'}^*(\,\cdot\,)). \end{equation*} An element~$A\in D_2(\mathcal{E})$ is called a \emph{Hamiltonian structure} on~$\mathcal{E}$ if $[\![ A,A]\!]=0$. Two Hamiltonian operators $A_1$ and $A_2$ are said to be \emph{compatible} if $[\![ A_1,A_2]\!]=0$ (cf.~Section~\ref{sec:gjsis_jets:hamilt-form}). \begin{remark} Using this definition of a Hamiltonian structure, the Hamiltonian formalisms on jet spaces, including the Magri scheme, explained in Section~\ref{sec:gjsis_jets:hamilt-form}, can be extended straightforwardly to the case of equations described here. \end{remark} \begin{example} The Camassa-Holm equation~\cite{CamassaHolm:ISWEPS} \begin{equation*} u_t-u_{txx}-uu_{xxx}-2u_xu_{xx}+3uu_x=0 \end{equation*} has a bi-Hamiltonian structure: \begin{equation*} A_1=D_x,\qquad A_2=-D_t-uD_x+u_x. \end{equation*} This equation is often written in the form \begin{align*} &m_t+um_x+2u_xm=0, \\ &m-u+u_{xx}=0. \end{align*} Then the bi-Hamiltonian structure takes the form \begin{equation*} A'_1= \begin{pmatrix} D_x & 0 \\ D_x-D_x^3 & 0 \end{pmatrix},\qquad A'_2= \begin{pmatrix} 0 & -1 \\ 2mD_x+m_x & 0 \end{pmatrix}. \end{equation*} \end{example} \begin{example} Let $\mathcal{E}$ be a bi-Hamiltonian equation given by $F=0$ and $A_1$ and $A_2$ be the corresponding Hamiltonian operators. The Kupershmidt deformation~\cite{Kupershmidt:KInS,KerstenKrasilshchikVerbovetskyVitolo:InKD}~$\tilde{\mathcal{E}}$ of~$\mathcal{E}$ has the form \begin{equation*} F+A_1^*(w)=0,\qquad A_2^*(w)=0, \end{equation*} where $w=(w^1,\dots,w^l)$ are new dependent variables. For example, the KdV6 equation~\cite{Karasu(Kalkani)KarasuSakovichSakovichTurhan:ANIGKdV} \begin{equation*} v_t+v_{xxx}+12vv_x-w_x=0,\qquad w_{xxx}+8vw_x+4wv_x=0, \end{equation*} is a Kupershmidt deformation of the KdV equation corresponding to the Hamiltonian operators~$A_1=D_x$ and~$A_2=D_x^3+8vD_x+4v_x$. The following two variational bivectors define a bi-Hamiltonian structures on~$\tilde{\mathcal{E}}$: \begin{equation*} \tilde{A}_1= \begin{pmatrix} A_1 & -A_1 \\ 0 & \ell_{F+A_1^*(w)+A_2^*(w)} \end{pmatrix}, \qquad \tilde{A}_2= \begin{pmatrix} A_2 & -A_2 \\ -\ell_{F+A_1^*(w)+A_2^*(w)} & 0 \end{pmatrix}. \end{equation*} \end{example} \begin{example} The equation \begin{equation*} z_{yy} + (1/z)_{xx} +2 =0 \end{equation*} associated with an integrable class of Weingarten surfaces~\cite{BaranMarvan:InWSFC} is bi-Hamiltonian with operators $D_x^2$ and $2zD_{xy}-z_yD_x+z_xD_y$. \end{example} \section{Nonlocal theory} \label{sec:gjsis_nonloc:nonlocal-theory} Nonlocal phenomena in the theory of integrable systems are quite common. Here by \emph{nonlocality} we mean an extension of the initial system by new variables (fields) that are related to the old ones by differential relations. Perhaps, the simplest way to observe how nonlocal objects originate is to analyse the action of recursion operators on symmetries. \begin{example} \label{exmp:gjsis_nonloc:1} Consider recursion operator~\eqref{eq:gjsis_jets:53} $R=D_x^2+4u+2u_1D_x^{-1}$ that generates the higher KdV equations (it can be shown that successive application of~$R$ to the first symmetry~$\phi_1=u_1$ results in polynomial expressions in~$u$, $u_1,\dots,u_k,\dots$, see, e.g.,~\cite{Krasilshchik:SMPLSH,Sergyeyev:LSGNInTDROpNApFS}). When one applies the operator~$R$ to the first $(x,t)$-dependent symmetry \begin{equation*} \bar{\phi}_1=tu_1+\frac{1}{6} \end{equation*} (the Galilean boost), this will result in the scaling symmetry \begin{equation*} \bar{\phi}_3=tu_3+(6tu+\frac{1}{3}x)u_1+\frac{2}{3}u, \end{equation*} but application of the recursion operator to~$\bar{\phi}_3$ leads to an expression that contains the nonlocal term~$D_x^{-1}(u)$ which can not be expressed in the geometrical terms introduced above\footnote{Of course, the Lenard operator itself contains a nonlocal summand, but we can consider it just as a convenient reformulation of the Magri relation~\eqref{eq:gjsis_jets:60}.}. An apparent way to incorporate this nonlocal object into the initial geometric setting is to introduce a new variable, say~$w$, that is related with the old one by~$w_x=u$. This relation, due to the KdV equation, implies another one:~$w_t=3u^2+u_{xx}$ and thus we shall result in the system \begin{equation*} u_t=6uu_x+u_{xxx},\qquad w_x=u,\qquad w_t=3u^2+u_{xx}. \end{equation*} \end{example} A general geometric formulation of this construction was first introduced in~\cite{VinogradovKrasilshchik:MCHSNEvEqNS,KrasilshchikVinogradov:NTGDEqSCLBT} and below we shall give a concise exposition of the theory together with a number of applications. \subsection{Differential coverings} \label{sec:gjsis_nonloc:diff-cover} The notion of a covering was already used in Section~\ref{sec:gjsis_eqs:diff-equat} in the context of the tangent and cotangent coverings. Here we discuss it in more detail. Let~$\mathcal{E}\subsetJ^\infty(\pi)$, where~$\pi\colon E\to M$, $\dim M=n$, be an equation. Consider a locally trivial bundle~$\tau\colon\tilde{\mathcal{E}}\to\mathcal{E}$ and endow the manifold~$\tilde{\mathcal{E}}$ with an $n$-dimensional distribution~$\tilde{\mathcal{\mathcal{C}}}$ in such a way that \begin{enumerate} \item\label{item:gjsis_nonloc:1} $\tilde{\mathcal{C}}$ is integrable and \item\label{item:gjsis_nonloc:2} for any point~$\tilde{\theta}\in\tilde{\mathcal{E}}$ the restriction~$\eval{\mathinner{\!}\mathrm{d}\tau}_{\tilde{\mathcal{\mathcal{C}}}_{\tilde{\theta}}}$ is a one-to-one correspondence between~$\tilde{\mathcal{\mathcal{C}}}_{\tilde{\theta}}$ and the Cartan plane~$\mathcal{C}_{\tau(\tilde{\theta})}$. \end{enumerate} Then we say that~$\psi$ is endowed with the structure of a \emph{differential covering} (or simply a \emph{covering}, to be short) over~$\mathcal{E}$. \begin{coordinates} Consider a trivialization of the bundle~$\tau$ and let~$w^1,\dots,w^j,\dots$ be fibre-wise coordinates (the so-called \emph{nonlocal variables}). The number~$r$ of these coordinates is called the \emph{dimension} of~$\tau$. Let~$D_1,\dots,D_n$ be the total derivatives on~$\mathcal{E}$. By Property~(\ref{item:gjsis_nonloc:2}) in the definition of the distribution~$\tilde{\mathcal{\mathcal{C}}}$ there exist $\tau$-vertical vector fields~$X_1,\dots,X_n$ on~$\tilde{\mathcal{E}}$ such that the fields \begin{equation*} \tilde{D}_i=D_i+X_i,\qquad i=1,\dots,n, \end{equation*} lie in~$\tilde{\mathcal{\mathcal{C}}}$. Then Property~(\ref{item:gjsis_nonloc:1}) is equivalent to the system of equations \begin{equation} \label{eq:gjsis_nonloc:1} D_i(X_j)-D_j(X_i)+[X_i,X_j]=0,\qquad 1\le i<j\le n, \end{equation} where~$X_1,\dots,X_n$ are $\tau$-vertical fields and~$D_i(X_j)$ denotes the component-wise action. Since the vector fields~$X_i$ are $\tau$-vertical fields, they can be presented in the form \begin{equation*} X_i=\sum_{j=1}^rX_i^j\frac{\partial}{\partial w^j}, \end{equation*} where~$X_i^j$ are smooth functions on~$\tilde{\mathcal{E}}$, while~$\mathcal{E}$, as a manifold with distribution, is isomorphic to the infinite prolongation of the system of PDEs \begin{equation*} \frac{\partial w^j}{\partial x^i}=X_i^j, \qquad i=1,\dots,n,\quad j=1,\dots,r, \end{equation*} which extends the initial equation~$\mathcal{E}$ and is compatible over it due to~\eqref{eq:gjsis_nonloc:1}. This system is called the \emph{covering equation}. \end{coordinates} \begin{example} \label{exmp:gjsis_nonloc:2} Consider the one-dimensional covering over the KdV equation determined by \begin{equation*} \tilde{D}_x=D_x+u\frac{\partial}{\partial w},\qquad \tilde{D}_t=D_t+(3u^2+u_2)\frac{\partial}{\partial w}. \end{equation*} The covering equation in this case is \begin{equation*} \frac{\partial w}{\partial x}=u,\qquad \frac{\partial w}{\partial t}=3u^2+u_{xx} \end{equation*} and is isomorphic to the potential KdV equation~$w_t=3w_x^2+w_{xxx}$. Note the the relation between~$w$ and~$u$ may be expressed in the form~$w=\int u\mathinner{\!}\mathrm{d} x$, or~$w=D_x^{-1}u$ and thus this is exactly the nonlocality that arose in Example~\ref{exmp:gjsis_nonloc:1}. \end{example} \begin{example} \label{exmp:gjsis_nonloc:4} Let again the base equation be the KdV and the covering be described by the system \begin{equation} \label{eq:gjsis_nonloc:2} X=u+w^2+\lambda,\qquad T=u_2+2wu_1+2u^2+2(w^2-\lambda)u-4\lambda(w^2+1), \end{equation} where~$\lambda\in\mathbb{R}$. Actually,~\eqref{eq:gjsis_nonloc:2} determines a one-parameter family of covering structures in the trivial bundle~$\mathcal{E}\times\mathbb{R}\to\mathcal{E}$, but the covering equation is isomorphic to the modified KdV equation~$w_t=6w^2w_x+w_{xxx}$ for any~$\lambda$. Of course, the covering under consideration is a geometric realization of the Miura transformation~\cite{Miura:KVEqGIRExNT}. \end{example} \begin{remark} \label{rem:gjsis_nonloc:1} Note that any differential substitution~$u=\phi(x,w,\dots,w_I,\dots)$ is associated with a covering over the initial equation, though this covering may be infinite-dimensional. For example, such is the covering over the KdV equation~$u_t-6uu_x+u_{xxx}=0$ determined with the Hirota substitution \begin{equation*} u=-2\frac{\partial^2}{\partial x^2}\ln w \end{equation*} (see~\cite{Hirota:ExSKVEqMCS}). Nevertheless, in spite of the infinite dimension of this covering, the covering space is isomorphic to the fourth-order scalar equation \begin{equation*} ww_{xt}-w_tw_x+w_{xxxx}w-4w_{xxx}w_x+3w_{xx}^2=0 \end{equation*} in one unknown function. \end{remark} \begin{example} \label{exmp:gjsis_nonloc:6} Let~$P$ be the module of sections for some vector bundle~$\xi$ over~$\mathcal{E}$. Then the bundle~$j_\infty^h\colonJ_h^\infty(P)\to\mathcal{E}$ of horizontal jets is an infinite-dimensional covering over~$\mathcal{E}$. If~$v_K^l$ are adapted coordinates in~$J_h^\infty(P)$ then the total derivatives lifted to~$J_h^\infty(P)$ are of the form \begin{equation} \label{eq:gjsis_nonloc:6} \tilde{D}_i=D_i+\sum_{l,K}v_{Ki}^l\frac{\partial}{\partial v_K^l}. \end{equation} This construction is generalised in the next example. \end{example} \begin{example}[$\Delta$-coverings] \label{exmp:gjsis_nonloc:5} Let~$\mathcal{E}$ be an equation and consider a $\mathcal{C}$-differential operator~$\Delta\colon P\to Q$, where~$P$ and~$Q$ are modules of sections for some vector bundles~$\xi$ and~$\zeta$ over~$\mathcal{E}$. Let~$\Phi_\Delta\colonJ_h^\infty(P)\to J_h^\infty(Q)$ be the corresponding morphism of vector bundles (see Proposition~\ref{prop:gjsis_jets:2}). Then, under natural conditions of non-degeneracy,~$\tilde{\mathcal{E}}_\Delta=\ker\Phi_\Delta$ is a sub-bundle in~$\xi_\infty\colonJ_h^\infty(P)\to\mathcal{E}$ that carries a natural structure of a covering: the total derivatives in this covering are obtained by restriction of the operators~\eqref{eq:gjsis_nonloc:6} to~$\tilde{\mathcal{E}}_\Delta$. We call this covering the \emph{$\Delta$-covering} over~$\mathcal{E}$. If the operator~$\Delta$ is locally given in the matrix form~$\Delta=\Vert\sum_Kd_{\alpha\beta}^KD_K\Vert$ then the subspace~$\tilde{\mathcal{E}}_\Delta\subsetJ_h^\infty(\xi)$ is described by the relations \begin{equation*} \sum_{\alpha,K} d_{\alpha\beta}^Kv_K^\alpha=0 \end{equation*} and their prolongations. Obviously, the tangent and cotangent coverings are particular cases of this construction with~$\Delta=\ell_{\mathcal{E}}$ and~$\Delta=\ell_{\mathcal{E}}^*$, respectively\footnote{Here and below we consider normal equations and use the notation~$\ell_{\mathcal{E}}$ instead of~$\bar{\ell}_F$ which is justified by the results of Subsection~\ref{sec:gjsis_eqs:line-diff-eqaut}}. $\Delta$-coverings play the key r\^{o}le in solving the following factorisation problem: let~$\Delta'\colon P'\to Q'$ be another $\mathcal{C}$-differential operator; how to find all operators~$A\colon P\to P'$ such that \begin{equation} \label{eq:gjsis_nonloc:7} \Delta'\circ A=B\circ\Delta, \end{equation} i.e., such that the diagram \begin{equation*} \begin{CD} P@>\Delta>>Q\\ @VAVV@VVBV\\ P'@>>\Delta'>Q' \end{CD} \end{equation*} is commutative? Note that any operator~$A$ of the form~$A=B'\circ\Delta$, where~$B'\colon Q\to P'$ is an arbitrary $\mathcal{C}$-differential operator, is a solution to~\eqref{eq:gjsis_nonloc:7}. Such solutions will be called \emph{trivial}. To find nontrivial solutions, first note that since~$\Delta'$ is a $\mathcal{C}$-differential operator it can be lifted to the covering just by changing the total derivatives~$D_i$ to the lifted ones~$\tilde{D}_i$. Denote this lift by~$\tilde{\Delta}'$. Second, let us put into correspondence to any operator~$A=\Vert\sum_Ka_{\alpha\beta}^K\Vert$ the vector-function \begin{equation*} \tilde{\Phi}_A= \eval{\left(\sum_{\alpha,K}a_{\alpha,1}^Kv_K^\alpha, \dots,\sum_{\alpha,K}a_{\alpha,r'}^Kv_K^\alpha\right)}_{\tilde{\mathcal{E}}_\Delta}, \qquad r'=\dim P', \end{equation*} Then one has: \begin{proposition} \label{prop:gjsis_nonloc:1} Classes of solutions of Equation~\eqref{eq:gjsis_nonloc:7} modulo trivial ones are in one-to-one correspondence with solutions of the equation \begin{equation*} \tilde{\Delta}'(\tilde{\Phi}_A)=0. \end{equation*} \end{proposition} Operators satisfying~\eqref{eq:gjsis_nonloc:7} take elements of~$\ker\Delta$ to those of~$\ker\Delta'$. \end{example} Consider system~\eqref{eq:gjsis_nonloc:1} that determines a covering structure in the space~$\tilde{\mathcal{E}}$ and assume that the coefficients~$X_i^j$ of the vertical vector fields~$X_i$ are independent of nonlocal variables~$w^\alpha$. In this case,~\eqref{eq:gjsis_nonloc:1} reduces to \begin{equation} \label{eq:gjsis_nonloc:3} D_i(X_j)=D_j(X_i),\qquad 1\le i<j\le n; \end{equation} the corresponding covering is called \emph{Abelian}. The covering in Example~\ref{exmp:gjsis_nonloc:2} is an Abelian one, while the covering associated with the Miura transformation (Example~\ref{exmp:gjsis_nonloc:4}) is not. Let~$\dim\tau=1$ and define a differential horizontal $1$-form on~$\mathcal{E}$ by setting \begin{equation} \label{eq:gjsis_nonloc:4} \omega_\tau=\sum_{i=1}^nX_i\mathinner{\!}\mathrm{d} x^i. \end{equation} Then~\eqref{eq:gjsis_nonloc:3} amounts to the equation \begin{equation} \label{eq:gjsis_nonloc:5} \rmd_h\omega_\tau=0, \end{equation} where~$\rmd_h$ is the horizontal differential on~$\mathcal{E}$. Thus, one-dimensional Abelian coverings over~$\mathcal{E}$ are in one-to-one correspondence with closed horizontal $(n-1)$-forms. In Example~\ref{exmp:gjsis_nonloc:4}, we presented a one-parameter family of coverings over the KdV equation. Are these coverings different for different values of the parameter~$\lambda$ or not and what the word ``different'' means in this context? The answer is the following. Consider an equation~$\mathcal{E}$ and two coverings~$\tau_i\colon\tilde{\mathcal{E}}_i\to\mathcal{E}$, $i=1$, $2$, over~$\mathcal{E}$. We say that these coverings are \emph{gauge equivalent} (or simply \emph{equivalent}) if there exists an isomorphism~$\phi\colon\tilde{\mathcal{E}}_1\to\tilde{\mathcal{E}}_2$ of the equations~$\tilde{\mathcal{E}}_1$ and~$\tilde{\mathcal{E}}_2$ such that the diagram \begin{equation*} \xymatrix{\tilde{\mathcal{E}}_1\ar[rr]^\phi\ar[rd]_{\tau_1}& &\ar[ld]^{\tau_2}\tilde{\mathcal{E}}_2\\ &\mathcal{E}& } \end{equation*} is commutative, i.e.,~$\tau_2\circ\phi=\tau_1$. In this sense, all coverings~\eqref{eq:gjsis_nonloc:2} are different, i.e., pair-wise non-equivalent for different values of~$\lambda$. A general cohomological technique to check whether a parameter can be eliminated or not was suggested in~\cite{Marvan:HGCNSP,Marvan:SPP}. We say that a covering~$\tau\colon\tilde{\mathcal{E}}\to\mathcal{E}$ is \emph{trivial} if for any point~$\theta\in\mathcal{E}$ there exists a neighbourhood~$\mathcal{U}\ni\theta$ such that \begin{enumerate} \item $\eval{\tau}_{\mathcal{U}}$ is a trivial bundle; \item there exists an adapted coordinate system in~$\mathcal{U}$ for which the fields~$\tilde{D}_i$ are of the form~$\tilde{D}_i=D_i$, $i=1,\dots,\dim M$. \end{enumerate} \begin{theorem} \label{thm:gjsis_nonloc:1} There exists a one-to-one correspondence between equivalence classes of one-dimensional Abelian coverings over~$\mathcal{E}$ and elements of the horizontal cohomology group~$H_h^1(\mathcal{E})$ given by~\eqref{eq:gjsis_nonloc:4}. In particular, a covering~$\tau$ is trivial if and only if the form~$\omega_\tau$ is a co-boundary, i.e.,~$\omega_\tau=\rmd_h(f)$. \end{theorem} Since~$H_h^1(\mathcal{E})$ coincides with the group of conservation laws when~$\dim M=2$, Theorem~\ref{thm:gjsis_nonloc:1} allows one to construct special type of coverings by conservation laws of the equation at hand. \begin{example} \label{exmp:gjsis_nonloc:3} The Camassa-Holm equation \begin{equation*} u_t-u_{txx}+3uu_x=2u_xu_{xx}+uu_{xxx} \end{equation*} admits the conservation law \begin{equation*} \omega=(u-u_{xx})\mathinner{\!}\mathrm{d} x+\frac{1}{2}(u_x^2-3u^2+2uu_{xx})\mathinner{\!}\mathrm{d} t; \end{equation*} consequently, the corresponding covering is given by \begin{equation*} w_x=u-u_{xx},\qquad w_t=\frac{1}{2}(u_x^2-3u^2+2uu_{xx}). \end{equation*} \end{example} \begin{remark} Since the group~$H_h^1(\mathcal{E})$ is trivial for normal equations in the case~$\dim M>2$, this implies that such equations possess no nontrivial one-dimensional Abelian covering. Actually, there exist very strong indications that these equations do not have finite-dimensional coverings at all (see~\cite{Marvan:SSC}). \end{remark} \begin{example}[the KP equation] Consider the dispersionless Kadomtsev-Petviashvili equation \begin{equation*} (u_t-6uu_x+u_{xxx})_x=u_{yy}. \end{equation*} It admits an obvious covering \begin{equation} \label{eq:gjsis_nonloc:8} w_x=u_y,\qquad w_y=u_t-6uu_x+u_{xxx}, \end{equation} which at first glance seems to be one-dimensional. But this is not the case, because Equations~\eqref{eq:gjsis_nonloc:8} do not contain information on the derivative~$w_t$. To incorporate these data, we must introduce infinite number of nonlocal variables~$w^0$, $w^1,\dots$ such that \begin{equation*} w^0=w,\ w_t^0=w^1,\,\dots,\,w_t^r=w^{r+1},\,\dots \end{equation*} and express their $x$- and $y$-derivatives using~\eqref{eq:gjsis_nonloc:8}. Thus, the covering is infinite-dimensional actually. \end{example} It was shown above that one-dimensional Abelian coverings can be constructed using conservation laws of the equation. Another type of coverings is related to Wahlquist-Estabrook prolongation structures~\cite{WahlquistEstabrook:PSNEvEq,WahlquistEstabrook:PSNEvEqII,DoddFordy:PSQF} and their description is based on the following \emph{ansatz}: Let~$u_t=f(u,u_1,\dots,u_k)$ be a system of evolution equations, $u=(u^1,\dots,u^m)$, $f=(f^1,\dots,f^m)$ being vectors and~$u_i$ denoting the $i$th derivative with respect to~$x$. Let us look for coverings such that the coefficients of the fields~$X$ and~$T$ in \begin{equation*} \tilde{D}_x=D_x+X,\qquad\tilde{D}_t=D_t+T \end{equation*} depend on~$u$, $u_1,\dots,u_{k-1}$ and nonlocal variables only. Then, locally, the description of such coverings locally reduces to representations of a certain free Lie algebra (the so-called \emph{Wahlquist-Estabrook algebra}) in vector fields on the fibre~$W$ of the trivial bundle~$\tau\colon\mathcal{E}\times W\to\mathcal{E}$. \begin{example} \label{exmp:gjsis_nonloc:7} Consider the potential KdV equation \begin{equation} \label{eq:gjsis_nonloc:9} u_t=u_x^2+u_{xxx} \end{equation} and let us describe coverings~$\tilde{D}_x=D_x+X$, $\tilde{D}_t=D_t+T$ over~$\mathcal{E}$ such that the fields~$X$ and~$T$ depend on~$u$, $u_1$ and~$u_2$ only. Straightforward computations show that all these coverings are of the form \begin{align}\label{eq:gjsis_nonloc:10} X&=u^2\mathbf{a}+u\mathbf{b}+\mathbf{c},\nonumber\\ T&=(2uu_2-u_1^2+2u^2u_1)\mathbf{a}+(u_2+2uu_1)\mathbf{b} +u_1[\mathbf{c},\mathbf{b}]+\frac{1}{2}u^2[\mathbf{b},\mathbf{d}] +u[\mathbf{c},\mathbf{d}]+\mathbf{e}, \end{align} where~$\mathbf{a}$, $\mathbf{b}$, $\mathbf{c}$, $\mathbf{d}$ and~$\mathbf{e}$ are vector fields on the fibre~$W$ of the covering (i.e., such that they do not depend on the equation coordinates) which enjoy the commutator relations \begin{gather*} 2\mathbf{a}=[\mathbf{a},\mathbf{b}],\quad \mathbf{b}=[\mathbf{a},\mathbf{c}],\quad \mathbf{d}=2\mathbf{c}+[\mathbf{c},\mathbf{b}],\\[0pt] [\mathbf{a},\mathbf{d}]=[\mathbf{c},\mathbf{e}]=0,\\[0pt] [\mathbf{b},\mathbf{d}]+\frac{1}{2}[\mathbf{b},[\mathbf{b},\mathbf{d}]]=0, \quad [\mathbf{b},\mathbf{e}]+[\mathbf{c},[\mathbf{c},\mathbf{d}]]=0,\\[0pt] [\mathbf{a},\mathbf{e}]+[\mathbf{b},[\mathbf{c},\mathbf{d}]] +\frac{1}{2}[\mathbf{c},[\mathbf{b},\mathbf{d}]]=0. \end{gather*} Now, to find all Wahlquist-Estabrook coverings for~\eqref{eq:gjsis_nonloc:9} amounts to describing representations, as vector fields on~$W$, of the free Lie algebra generated by the elements~$\mathbf{a}$, $\mathbf{b}$, $\mathbf{c}$, $\mathbf{d}$, and~$\mathbf{e}$ with the above relations. If~$W=\mathbb{R}$ then all such representations, up to an isomorphism, are \begin{gather*} \mathbf{a}\mapsto\frac{\partial}{\partial w},\quad \mathbf{b}\mapsto(2w+\beta)\frac{\partial}{\partial w},\quad \mathbf{c}\mapsto(w^2+\beta w+\gamma)\frac{\partial}{\partial w},\\ \mathbf{d}\mapsto-\Delta\frac{\partial}{\partial w},\quad \mathbf{e}\mapsto\Delta(w^2+\beta w+\gamma)\frac{\partial}{\partial w}, \end{gather*} where~$\beta$, $\gamma\in\mathbb{R}$ and~$\Delta=\beta^2-4\gamma$. The corresponding one-dimensional coverings, up to gauge equivalence, are of the form \begin{equation*} X=(u^2+2wu+w^2+\gamma)\frac{\partial}{\partial w} \end{equation*} (the parameter~$\beta$ can be removed by a gauge transformation) and with~$T$ given by~\eqref{eq:gjsis_nonloc:10}; they are pair-wise inequivalent for different values of~$\gamma$. \end{example} \begin{remark} The term \emph{covering} also refers to the parallel between classical differential geometry and geometry of PDEs. Namely, if we define dimension of an equation~$\mathcal{E}$ (or of a jet space~$J^\infty(\pi)$) as that of the corresponding Cartan distribution (i.e., the number of independent variables) then fibres of a differential covering become zero-dimensional, and this complies with the definition of a topological covering. That was the initial reason to name the object in~\cite{VinogradovKrasilshchik:MCHSNEvEqNS}. But the parallel goes far beyond this trivial observation. In~\cite{Igonin:AnCFGCPDEq}, a new powerful invariant (the \emph{fundamental Lie algebra}) of differential equations was proposed whose r\^{o}le in the theory of differential coverings is quite similar to the one that the fundamental group plays in topology. In particular, the fundamental Lie algebra allows one to enumerate (locally) all coverings over a given equation in the same way as conjugacy classes of subgroups of the fundamental group enumerate topological coverings. So, the dictionary evolved in the previous sections can be continued: \begin{eqnarray*} \text{\textbf{Manifold~$M$}}&&\text{\textbf{Differential equation~$\mathcal{E}$}} \\[1ex] \text{topological dimension}&\quad\longleftrightarrow\quad&\text{differential dimension} \\ \text{topological coverings}&\quad\longleftrightarrow\quad&\text{differential coverings} \\ \text{fundamental group}&\quad\longleftrightarrow\quad&\text{fundamental Lie algebra} \end{eqnarray*} \end{remark} Note also that using the fundamental Lie algebra technique the author of~\cite{Igonin:AnCFGCPDEq} proved \emph{nonexistence} of B\"{a}cklund transformations for some pairs of differential equations. It seems that it is impossible to achieve such a result by other methods. \subsection{Nonlocal symmetries} \label{sec:gjsis_nonloc:nonlocal-symmetries} The concept of a symmetry discussed in Section~\ref{sec:gjsis_eqs:diff-equat} can be generalised to the nonlocal situation. Consider an example. \begin{example} \label{exmp:gjsis_nonloc:8} Let \begin{equation} \label{eq:gjsis_nonloc:11} u_t=uu_x+u_{xx} \end{equation} be the Burgers equation (its Lie algebra of symmetries was fully described in~\cite{VinogradovKrasilshchik:MCHSNEvEqNS}). Direct computations show that~\eqref{eq:gjsis_nonloc:11} does not possess symmetries of the form~$\phi=\phi(x,t,u)$, but if one extends the setting by a new (nonlocal) variable~$w$ such that \begin{equation} \label{eq:gjsis_nonloc:12} w_x=u,\qquad w_t=\frac{1}{2}u^2+u_x \end{equation} then the equation~$\ell_{\mathcal{E}}(\phi)=0$ will acquire a new family of solutions of the form \begin{equation} \label{eq:gjsis_nonloc:13} \phi=(au-2a_x)\mathrm{e}^{-\frac{1}{2}w}, \end{equation} where~$a=a(x,t)$ is an arbitrary solution of the heat equation~$a_t=a_{xx}$. \end{example} The question is: can functions~\eqref{eq:gjsis_nonloc:13} be considered as symmetries of Equation~\eqref{eq:gjsis_nonloc:11} in some natural sense? To answer this question, consider an arbitrary equation~$\mathcal{E}\subsetJ^\infty(\pi)$ and a covering~$\tau\colon\tilde{\mathcal{E}}\to\mathcal{E}$. We say that~$\phi$ is a \emph{nonlocal symmetry} (or $\tau$-symmetry) of~$\mathcal{E}$ if it is a symmetry of~$\tilde{\mathcal{E}}$. \begin{coordinates} Let~$\mathcal{E}\subsetJ^\infty(\pi)$ be an equation and~$\tau\colon\tilde{\mathcal{E}}\to\mathcal{E}$ be a covering locally given by the total derivatives \begin{equation*} \tilde{D}_i=D_i+\sum_jX_i^j\frac{\partial}{\partial w^j},\qquad i=1,\dots,\dim M. \end{equation*} Then any nonlocal $\tau$-symmetry is of the form \begin{equation} \label{eq:gjsis_nonloc:14} \tilde{\bi{E}}_\phi+\sum_j\psi^j\frac{\partial}{\partial w^j}. \end{equation} Here~$\phi$ is an $m$-component vector-function on~$\tilde{\mathcal{E}}$ that satisfies the equation \begin{equation}\label{eq:gjsis_nonloc:15} \tilde{\ell}_{\mathcal{E}}(\phi)=0, \end{equation} $\psi^j$ are functions on~$\tilde{\mathcal{E}}$ such that \begin{equation} \label{eq:gjsis_nonloc:16} \tilde{D}_i(\psi^j)=\tilde{\ell}_{X_i^j}(\phi) +\sum_\alpha\frac{\partial X_i^j}{\partial w^\alpha}\psi^\alpha \end{equation} and \begin{equation} \label{eq:gjsis_nonloc:17} \tilde{\bi{E}}_\phi= \sum_{\substack{\text{over internal} \\ \text{coordinates}}} \tilde{D}_I(\phi^j)\frac{\partial}{\partial u_I^j} \end{equation} (recall that the ``tilde'' over a $\mathcal{C}$-differential operator denotes its natural lifting to the covering). \end{coordinates} \begin{example} \label{exmp:gjsis_nonloc:9} Let us consider Example~\ref{exmp:gjsis_nonloc:8} again. In the case of covering~\eqref{eq:gjsis_nonloc:12} Equations~\eqref{eq:gjsis_nonloc:16} take the form \begin{equation} \label{eq:gjsis_nonloc:18} \tilde{D}_x(\psi)=\phi,\qquad \tilde{D}_t(\psi)=u\phi+\tilde{D}_x(\phi). \end{equation} Consequently, for~$\phi$ of the form~\eqref{eq:gjsis_nonloc:13} we see that \begin{equation*} \psi=-2a\mathrm{e}^{-\frac{1}{2}w} \end{equation*} satisfies~\eqref{eq:gjsis_nonloc:18}. Thus, the pair of functions~$\phi$ and~$\psi$ determine a nonlocal symmetry of the Burgers equation in the sense of the above definition. \end{example} However, the situation of the previous example is not generic. \begin{example} \label{exmp:gjsis_nonloc:10} Consider the covering \begin{equation*} w_x=u,\qquad w_t=3u^2+u_{xx} \end{equation*} over the KdV equation~$u_t=6uu_x+u_{xxx}$ and let us try to find nonlocal symmetries in this covering. In the case under consideration, Equations~\eqref{eq:gjsis_nonloc:16} acquire the form \begin{equation} \label{eq:gjsis_nonloc:19} \tilde{D}_x(\psi)=\phi,\qquad\tilde{D}_t(\psi)=6u\phi+\tilde{D}_x^2(\phi), \end{equation} while~\eqref{eq:gjsis_nonloc:15} is \begin{equation*} \tilde{D}_t(\phi)=6u_1\phi+6u\tilde{D}_x(\phi)+\tilde{D}_x^3(\phi). \end{equation*} The simplest solution of the last equation that depends on~$w$ is \begin{equation*} \phi=tu_5+\left(10tu+\frac{1}{3}x\right)u_3+ 4\left(5tu_1+\frac{1}{3}\right)u_2+ 2\left(15tu^2+xu+\frac{1}{3}w\right)u_1+\frac{8}{3}u^2. \end{equation*} But solving~\eqref{eq:gjsis_nonloc:19} with~$\phi$ of the above form leads to a contradiction: no function~$\psi$ exists on~$\mathcal{E}$ such that~\eqref{eq:gjsis_nonloc:19} is valid for our~$\phi$. Nevertheless, if we introduce another nonlocal variable~$w'$ satisfying \begin{equation*} w_x'=u^2,\qquad w_t'=4u^3-u_x^2+2uu_{xx} \end{equation*} then~\eqref{eq:gjsis_nonloc:19} will be resolved in the new setting. But a similar problem arises at the next step: now we need to reconstruct the coefficient~$\psi'$ at~$\partial/\partial w'$. \end{example} The procedure we encountered in Example~\ref{exmp:gjsis_nonloc:10} is typical and we shall describe it in general terms now. Let~$\tau\colon\tilde{\mathcal{E}}\to\mathcal{E}$ be a covering and denote by~$\mathcal{F}$ and~$\tilde{\mathcal{F}}$ the function algebras on~$\mathcal{E}$ and~$\tilde{\mathcal{E}}$, respectively. An $\mathbb{R}$-linear map~$X\colon\mathcal{F}\to\tilde{\mathcal{F}}$ is called a \emph{$\tau$-shadow} if \begin{enumerate} \item $X$ is a derivation, i.e., \begin{equation*} X(fg)=fX(g)+gX(f) \end{equation*} for all~$f$, $g\in\mathcal{F}$; \item the action of~$X$ preserves the Cartan distribution, i.e.,~$\mathrm{L}_X(\omega)\in\La_{\scriptscriptstyle{\mathcal{C}}}(\tilde{\mathcal{E}})$ as soon as~$\omega\in\La_{\scriptscriptstyle{\mathcal{C}}}(\mathcal{E})$ (or, equivalently, for any Cartan field~$\tilde{Y}$ on~$\tilde{\mathcal{E}}$ and its restriction~$Y$ to~$\mathcal{F}$ the commutator~$[X,\tilde{Y}]=XY-\tilde{Y}X$ is a Cartan field again). \end{enumerate} In particular, any symmetry of~$\tilde{\mathcal{E}}$ can be considered as a shadow in an arbitrary covering~$\tau$. \begin{coordinates} Let~$\mathcal{U}$ be the set of internal coordinates on~$\mathcal{E}$. Then any $\tau$-shadow is given by the formula \begin{equation*} \tilde{\bi{E}}_\phi= \sum_{u_I^j\in\mathcal{U}}\tilde{D}_I(\phi^j)\frac{\partial}{\partial u_I^j}, \end{equation*} where~$\phi^1,\dots,\phi^m$ are functions on~$\tilde{\mathcal{E}}$ and~$\tilde{D}_1,\dots,\tilde{D}_n$ are total derivatives on~$\tilde{\mathcal{E}}$ (cf.~\eqref{eq:gjsis_nonloc:17}). The vector function~$\phi=(\phi^1,\dots,\phi^m)$ must satisfy the equation \begin{equation*} \tilde{\ell}_{\mathcal{E}}(\phi)=0. \end{equation*} \end{coordinates} We say that a $\tau$-shadow~$X$ is \emph{reconstructed} in~$\tau$ if there exists a nonlocal $\tau$-symmetry~$\tilde{X}$ such that~$\eval{\tilde{X}}_{\mathcal{F}}=X$. As Examples~\ref{exmp:gjsis_nonloc:9} and~\ref{exmp:gjsis_nonloc:10} show, not all shadows can be reconstructed in a straightforward way. A general result that describes the reconstruction procedure was proved in~\cite{Khorkova:CLNS} (see also~\cite{KrasilshchikVinogradov:NTGDEqSCLBT}): \begin{proposition} \label{prop:gjsis_nonloc:2} Let~$\tau\colon\tilde{\mathcal{E}}\to\mathcal{E}$ be a covering and~$X$ be a $\tau$-shadow. Then there exists another covering~$\bar{\tau}\colon\bar{\mathcal{E}}\to\tilde{\mathcal{E}}$ and a $\bar{\tau}$-shadow~$\bar{X}$ such that~$\eval{\bar{X}}_{\mathcal{F}}=X$. \end{proposition} Thus, putting~$\tau=\tau_0$ and~$\tau_{i+1}=\tilde{\tau}_i$ and applying Proposition~\ref{prop:gjsis_nonloc:2} sufficiently (maybe, infinitely) many times we shall arrive to a covering in which the given shadow is reconstructed. \begin{coordinates} Actually, the results of~\cite{Khorkova:CLNS} not just state the existence of the needed covering but provide a canonical way to construct the one. The construction is in a sense tautological and mimics relations~\eqref{eq:gjsis_nonloc:16}. \begin{proposition} \label{prop:gjsis_nonloc:3} Let~$\tau\colon\tilde{\mathcal{E}}\to\mathcal{E}$ be a covering over an equation~$\mathcal{E}$ with nonlocal coordinates~$w^1,\dots,w^j,\dots$ and~$\tilde{D}_1,\dots,\tilde{D}_n$ be total derivatives in this covering. Let also~$\phi$ be a $\tau$-shadow. Then: \begin{enumerate} \item the relations \begin{equation} \label{eq:gjsis_nonloc:20} \frac{\partial\tilde{w}^j}{\partial x^i}=\tilde{\ell}_{X_i^j}(\phi) +\sum_\alpha\frac{\partial X_i^j}{\partial w^\alpha}\tilde{w}^\alpha,\qquad i=1,\dots,n,\quad j=1,\dots,\dim\tau, \end{equation} define a covering over~$\tilde{\mathcal{E}}$ whose dimension equals that of~$\tau$; \item equations~\eqref{eq:gjsis_nonloc:16} are solvable in this covering. \end{enumerate} \end{proposition} \end{coordinates} \begin{remark} It follows from~\eqref{eq:gjsis_nonloc:20} that for an Abelian covering~$\tau$ the covering~$\tilde{\tau}$ is Abelian as well. Hence, at every step of reconstruction obstructions to solving~\eqref{eq:gjsis_nonloc:16} lie in the horizontal cohomology group of the corresponding equation. Consequently, if~$\tau\colon\tilde{\mathcal{E}}\to\mathcal{E}$ is an Abelian covering and~$H_h^1(\tilde{\mathcal{E}})=0$ then any $\tau$-shadow can be reconstructed to a nonlocal $\tau$-symmetry. In particular, any local symmetry of~$\mathcal{E}$ can be lifted to~$\tilde{\mathcal{E}}$. Let~$\mathcal{E}$ be an equation and~$\{[\omega^\alpha]\}$, $\omega^\alpha\in\La_h^1(\mathcal{E})$, be an $\mathbb{R}$-basis of the group~$H_h^1(\mathcal{E})$. Assume that \begin{equation*} \omega^\alpha=X_1^\alpha\mathinner{\!}\mathrm{d} x^1+\dots+X_n^\alpha\mathinner{\!}\mathrm{d} x^n \end{equation*} and consider the covering~$\tau_1\colon\mathcal{E}_1\to\mathcal{E}$ determined by \begin{equation*} \frac{\partial w^\alpha}{\partial x^i}=X_i^\alpha \end{equation*} for all~$\alpha$ and~$i=1,\dots,n$. For~$\mathcal{E}_1$ let us construct the covering~$\tau_2\colon\mathcal{E}_2\to\mathcal{E}_1$ in a similar way, etc. The covering~$\tau_*\colon\mathcal{E}_*\to\mathcal{E}$ obtained as the inverse limit of the sequence \begin{equation*} \dots\xrightarrow{\tau_{i+1}}\mathcal{E}_i\xrightarrow{\tau_i} \mathcal{E}_{i-1} \xrightarrow{\tau_{i-1}}\dots\xrightarrow{\tau_2}\mathcal{E}_1 \xrightarrow{\tau_1}\mathcal{E} \end{equation*} is called the \emph{universal Abelian covering} over~$\mathcal{E}$. By construction,~$H_h^1(\mathcal{E}_*)=0$. \begin{proposition} For an arbitrary Abelian covering~$\tau\colon\tilde{\mathcal{E}}\to\mathcal{E}$, there exists a uniquely (up to a gauge equivalence) defined morphism \begin{equation*} \xymatrix{ \mathcal{E}_*\ar[rr]^{f}\ar[dr]_{\tau_*}&& \tilde{\mathcal{E}}\ar[dl]^{\tau}\\ &\mathcal{E}& } \end{equation*} and any $\tau$-shadow can be reconstructed in~$\tau_*$. In particular, any symmetry of~$\mathcal{E}$ can be lifted to~$\mathcal{E}_*$. \end{proposition} \end{remark} \begin{remark} Though the covering~$\tilde{\tau}$ whose existence is stated in Proposition~\ref{prop:gjsis_nonloc:2} is determined canonically by the shadow~$X$, the new shadow~$\tilde{X}$ is not unique, but is defined up to an infinitesimal gauge symmetries of~$\tau$, i.e., up to~$Y\in\sym(\tilde{\mathcal{E}})$ such that~$\eval{Y}_{\mathcal{F}}=0$. Due to~\eqref{eq:gjsis_nonloc:16}, these symmetries are given by the equations \begin{equation} \label{eq:gjsis_nonloc:21} \tilde{D}_i(\psi^j)= \sum_\alpha\frac{\partial X_i^j}{\partial w^\alpha}\psi^\alpha \end{equation} and are of the form~$Y=\sum_\alpha\psi^\alpha\partial/\partial w^\alpha$. \begin{example} For Example~\ref{exmp:gjsis_nonloc:2}, Equations~\eqref{eq:gjsis_nonloc:21} take the form \begin{equation*} \tilde{D}_x(\psi)=0,\qquad \tilde{D}_t(\psi)=0 \end{equation*} and consequently infinitesimal gauge symmetries are~$\gamma\partial/\partial w$ in this case, $\gamma\in\mathbb{R}$. On the other hand, if we consider Example~\ref{exmp:gjsis_nonloc:4} then Equations~\eqref{eq:gjsis_nonloc:21} are written as \begin{equation*} \tilde{D}_x(\psi)=2w\psi,\qquad \tilde{D}_t(\psi)=2(u_1+2uw-4\lambda w)\psi. \end{equation*} The only solution of this system is~$\psi=0$ and thus there is no ambiguity in shadow reconstruction in this case. \end{example} Non-uniqueness of the solution to the problem of reconstruction leads, in turn, to the problem of commutation for shadows: no well defined way to compute the Lie bracket of shadows is known. This problem was first indicated in~\cite{OlverSandersWang:GS}. A way to solve it was suggested in~\cite{VerbovetskyGolovkoKrasilshchik:LBNS}, but a practical realization of the approach is somewhat cumbersome. \end{remark} To conclude this subsection, let us make a remark also related to the problem of reconstruction. Let~$\tau\colon\tilde{\mathcal{E}}\to\mathcal{E}$ be a finite-dimensional covering and~$X$ be a symmetry of~$\mathcal{E}$. One can (at least, locally) lift~$X$ to~$\tilde{\mathcal{E}}$ in an arbitrary way. Then, if~$X$ is an integrable vector field (i.e., if it possesses the corresponding one-parameter group of transformations) then the lifted field~$\tilde{X}$ is integrable as well. If~$\tilde{X}$ is a symmetry of~$\tilde{\mathcal{E}}$ then this means that we managed to reconstruct~$X$ up to a nonlocal symmetry in the covering~$\tau$. Conversely, consider the one-parameter group of transformations~$\{\tilde{A}_\lambda\}$ corresponding to~$\tilde{X}$ and for any~$\lambda\in\mathbb{R}$ define an $n$-dimensional distribution~$\tilde{\mathcal{C}}^\lambda$ on~$\tilde{\mathcal{E}}$ by \begin{equation} \label{eq:gjsis_nonloc:22} \tilde{\mathcal{C}}^\lambda\colon\theta\mapsto\tilde{\mathcal{C}}_\theta^\lambda= \tilde{A}_{\lambda,*}\left(\tilde{\mathcal{C}}_{\tilde{A}_\lambda^{-1}(\theta)}\right) \end{equation} where~$\theta\in\tilde{\mathcal{E}}$, $\tilde{\mathcal{C}}_\theta$ is the Cartan plane at the point~$\theta$ and~$F_*$ denotes the differential of the map~$F$. \begin{proposition} Correspondence~\eqref{eq:gjsis_nonloc:22} determines a one-parameter family~$\tau_\lambda$ of pair-wise inequivalent coverings over~$\tilde{\mathcal{E}}$ such that~$\tau_0=\tau$. \end{proposition} \begin{example} Take the covering \begin{equation*} X=u+w^2,\qquad T=u_2+2wu_1+2u^2+2w^2u \end{equation*} over the KdV equation and apply Proposition~\ref{prop:gjsis_nonloc:3} using the Galilean boost $tu_1+1/6$ for the symmetry~$X$. This will result in the Miura covering described in Example~\ref{exmp:gjsis_nonloc:4}. \end{example} Not all one-parameter families of coverings can be obtained by this procedure (a counter-example can be found in~\cite{Cieslinski:NSWAlIsInG,Cieslinski:GInSPCNNSS}). But a weaker result was proved in~\cite{IgoninKrasilshchik:OnPFBT}: \begin{theorem} Let~$\tau_\lambda\colon\tilde{\mathcal{E}}\to\mathcal{E}$ be a one-parameter family of coverings regarded as a deformation of the covering~$\tau=\tau_0$. Then the corresponding infinitesimal deformation is a $\tau$-shadow. \end{theorem} \subsection{B\"{a}cklund transformations and zero-curvature representations} \label{sec:gjsis_nonloc:zero-curv-repr} In conclusion, let us briefly discuss how the constructions of B\"{a}cklund transformations~\cite{RogersShadwick:BTTAp,RogersSchief:BDT} and zero-curvature representations~\cite{AblowitzKaupNewellSegur:InSTFAnNP} are translated to the geometrical language. A \emph{B\"{a}cklund transformation} between two equations~$\mathcal{E}'$ and~$\mathcal{E}''$ with the unknown functions~$u'$ and~$u''$, respectively, is another equation~$\mathcal{E}$ in unknown functions both~$u'$ and~$u''$ such that for any solution~$u'$ of~$\mathcal{E}'$ a solution~$u''$ of~$\mathcal{E}$ is a solution of~$\mathcal{E}''$ as well and vice versa. If~$\mathcal{E}'$ coincides with~$\mathcal{E}''$ then one speaks about \emph{auto-B\"{a}cklund transformation}. \begin{example} \label{exmp:gjsis_nonloc:11} Consider the sine-Gordon equation \begin{equation} \label{eq:gjsis_nonloc:23} u_{xy}=\sin u. \end{equation} Then the system \begin{equation} \label{eq:gjsis_nonloc:24} v_y-u_y=2\lambda\sin\frac{v+u}{2},\qquad v_x+u_x=\frac{2}{\lambda}\sin\frac{v-u}{2}, \end{equation} where~$\lambda\neq 0$ is a real parameter, determines the classical auto-B\"{a}cklund transformation (a one-parameter family, actually) for~\eqref{eq:gjsis_nonloc:23}, see~\cite{DoddBullough:BTSGEq}. \end{example} \begin{example} \label{exmp:gjsis_nonloc:12} The second example (which now can also be considered as a classical one) was found in~\cite{WahlquistEstabrook:PSNEvEq}. It is of the form \begin{align} \label{eq:gjsis_nonloc:25} &\left(\frac{v+w}{2}\right)_x +\left(\frac{v-w}{2}\right)^2+\lambda^2=0, \quad\lambda\in\mathbb{R},\nonumber\\ &\left(\frac{v-w}{2}\right)_t+ 6\left(\frac{v+w}{2}\right)_x\left(\frac{v-w}{2}\right)_x+ \left(\frac{v-w}{2}\right)_{xxx}=0 \end{align} and relates solutions of the KdV equation to each other (or, to be more precise, system~\eqref{eq:gjsis_nonloc:25} is a B\"{a}cklund transformation for the potential KdV equation, while solutions of the KdV itself are obtained by~$u=v_x$). \end{example} Analysis of these two examples (as well as other ones) shows that a B\"{a}cklund transformation between equations~$\mathcal{E}'$ and~$\mathcal{E}''$ is a diagram \begin{equation*} \xymatrix{ &\mathcal{E}\ar[dl]_{\tau'}\ar[dr]^{\tau''}&\\ \mathcal{E}'&&\mathcal{E}''\rlap{,} } \end{equation*} where~$\tau'$ and~$\tau''$ are coverings. The correspondence between solutions of~$\mathcal{E}'$ and~$\mathcal{E}''$ is achieved in the following way. Let~$u'=u'(x')$ be a solution of~$\mathcal{E}'$ and assume that~$\tau'$ is a finite-dimensional covering. Then the Cartan distribution of the equation~$\mathcal{E}$ induces on the finite-dimensional manifold~$\mathcal{E}_{u'}=(\tau')^{-1}(u')\subset\mathcal{E}$ an $n$-dimensional integrable distribution. In the vicinity of a generic point the latter possesses a $(\dim\tau')$-parameter family of maximal integral manifolds that are projected to~$u'$ by~$\tau'$ and to the corresponding family of solutions of~$\mathcal{E}''$ by~$\tau''$. Generically, such a correspondence is non-trivial provided~$\tau'$ and~$\tau''$ are not gauge equivalent. \begin{example} \label{exmp:gjsis_nonloc:13} A common way to construct non-trivial B\"{a}cklund transformations is the following. Let~$\tau\colon\tilde{\mathcal{E}}\to\mathcal{E}$ be a covering and~$f\colon\tilde{\mathcal{E}}\to\tilde{\mathcal{E}}$ be a finite symmetry of~$\tilde{\mathcal{E}}$, i.e., a diffeomorphism preserving the Cartan distribution. Then the composition~$\tau'=\tau\circ f$ is a covering as well and the pair~$(\tau,\tau')$ is an auto-B\"{a}cklund transformation for~$\mathcal{E}$. If~$f$ is not a gauge equivalence then this transformation is non-trivial. Consider covering~\eqref{eq:gjsis_nonloc:2} from Example~\ref{exmp:gjsis_nonloc:4} and note that the change of the nonlocal variable~$w\leftrightarrow-w$ is a symmetry of the covering equation (the mKdV one), but is not a gauge symmetry of the covering itself. Thus, for any value of the parameter~$\lambda$ we get an auto-B\"{a}cklund transformation of the KdV equation, i.e., a one-parameter family of B\"{a}cklund transformations. Note that the Wahlquist-Estabrook construction (Example~\ref{exmp:gjsis_nonloc:12}) is a consequence of the latter one. \end{example} \begin{remark} Families of B\"{a}cklund transformations, like the ones from Examples~\ref{exmp:gjsis_nonloc:11} and~\ref{exmp:gjsis_nonloc:12}, give one an opportunity to construct special exact solutions of integrable equations (such as multi-kink solutions for the sine-Gordon equation, multi-soliton solutions for the KdV, etc.). The construction uses the \emph{nonlinear superposition principle} which, in turn, is based on the following informal statement: \begin{theorem}[the Bianchi Permutability Theorem] Assume that an equation~$\mathcal{E}$ possesses a one-parameter family of auto-B\"{a}cklund transformations~$\mathcal{B}_\lambda$ and let~$\lambda\in\mathbb{R}$ be the parameter. For any solution~$u=u(x)$ of~$\mathcal{E}$ denote by~$\mathcal{B}_\lambda(u)$ the set of solutions obtained from~$u$ by means of~$\mathcal{B}_\lambda$. Then for any~$\lambda_1\neq\lambda_2$ there exists a solution~$u_{\lambda_1,\lambda_2}\in \mathcal{B}_{\lambda_1}(\mathcal{B}_{\lambda_2}(u))\cap \mathcal{B}_{\lambda_2}(\mathcal{B}_{\lambda_1}(u))$ that is expressed as a bi-differential operator applied to some solutions~$u_1\in\mathcal{B}_{\lambda_1}(u)$ and~$u_2\in\mathcal{B}_{\lambda_2}(u)$. \end{theorem} This ``theorem'' was first observed by Bianchi in~\cite{Bianchi:LGD} (see also~\cite{RogersShadwick:BTTAp}) in application to the sine-Gordon equation (Example~\ref{exmp:gjsis_nonloc:11}) and since then dozens of examples were computed, but nevertheless a general formulation of this statement (and, consequently, its general proof) is unknown to us. Some hints to a rigorous approach to the problem can be found in~\cite{Marvan:SLPBT}. \end{remark} Geometrical theory of B\"{a}cklund transformations is also related to an unorthodox approach to recursion operators~\cite{Marvan:AnLROp}. Consider an equation~$\mathcal{E}$ and its tangent covering~$\tau\colon\T(\mathcal{E})\to\mathcal{E}$ (see Section~\ref{sec:gjsis_eqs:diff-equat}). Recall that symmetries of~$\mathcal{E}$ are identified with sections of~$\tau$ that take the Cartan distribution on~$\mathcal{E}$ to that on~$\T(\mathcal{E})$. Hence, if we consider a diagram of the form \begin{equation*} \xymatrix{ &\tilde{\mathcal{E}}\ar[dl]_{\tau'}\ar[dr]^{\tau''}&\\ \T(\mathcal{E})\ar[dr]_{\tau}&&\T(\mathcal{E})\ar[dl]^{\tau}\\ &\mathcal{E}\rlap{,}& } \end{equation*} where~$\tau'$ and~$\tau''$ are coverings, then this B\"{a}cklund transformation will relate symmetries of~$\mathcal{E}$ to each other. Thus, this B\"{a}cklund transformation plays the r\^{o}le of a recursion operator for symmetries of~$\mathcal{E}$. \begin{example} \label{exmp:gjsis_nonloc:14} Consider the KdV equation~$u_t=6uu_x+u_{xxx}$ and two copies of its tangent covering with the new dependent variable~$v$ that enjoys the the additional equation \begin{equation*} v_t=6u_xv+6uv_x+v_{xxx}. \end{equation*} Introduce a nonlocal variable~$\tilde{v}$ by setting \begin{equation*} \tilde{v}_x=v,\qquad\tilde{v}_t=6uv+v_{xx}. \end{equation*} Thus, internal coordinates in~$\tilde{\mathcal{E}}$ are \begin{equation*} x,\ t,\ u=u_0,\ u_x=u_1,\,\dots,\,v=v_0,\ v_x=v_1,\,\dots,\,\tilde{v}. \end{equation*} Define the covering~$\tau'$ by \begin{equation*} \tau'\colon(x,t,u_k,v_k,\tilde{v})\mapsto(x,t,u_k,v_k) \end{equation*} and the covering~$\tau''$ by \begin{equation*} \tau'\colon(x,t,u_k,v_k,\tilde{v})\mapsto (x,t,u_k,D_x^k(v_2+4uv+2u_1\tilde{v})). \end{equation*} The B\"{a}cklund transformation obtained in such a way is the geometrical realization of the Lenard recursion operator~$R=D_x^2+4u+2u_1D_x^{-1}$ given in Equation~\eqref{eq:gjsis_jets:53}. \end{example} Another, less trivial example will be considered later (see Example~\ref{exmp:gjsis_nonloc:16} below), after discussing the concept of \emph{zero-curvature representations} (ZCR). Let~$\tau\colon\tilde{\mathcal{E}}\to\mathcal{E}$ be a covering. We say that it is \emph{linear} if \begin{enumerate} \item $\tau$ is a vector bundle; \item the action of vector fields~$\tilde{D}_1,\dots,\tilde{D}_n$ on~$\mathcal{F}(\tilde{\mathcal{E}})$ preserves the subspace of fibre-wise linear functions. \end{enumerate} \begin{coordinates} Let~$v^1,\dots,v^r,\dots$ be local coordinates along the fibre of~$\tau$ and the covering be given by the total derivatives \begin{equation} \label{eq:gjsis_nonloc:26} \tilde{D}_i=D_i+\sum_r X_i^r\frac{\partial}{\partial v^r},\qquad i=1,\dots,n. \end{equation} Then the covering is linear if and only if the coefficients~$X_i^r$ in~\eqref{eq:gjsis_nonloc:26} are of the form \begin{equation*} X_i^r=\sum_\alpha X_{i\alpha}^rv^\alpha, \end{equation*} where~$X_{i\alpha}^r$ are smooth functions on~$\mathcal{E}$. If we now identify the vertical terms~$X_i=\sum_r X_i^r\partial/\partial v^r$ in~\eqref{eq:gjsis_nonloc:26} with the function-valued matrices \begin{equation*} X_i= \begin{pmatrix} X_{i1}^1&\dots&X_{i1}^n\\ \dots&\dots&\dots\\ X_{in}^1&\dots&X_{in}^n \end{pmatrix} \end{equation*} then~\eqref{eq:gjsis_nonloc:26} will be rewritten as \begin{equation*} \tilde{D}_i=D_i+X_i,\qquad i=1,\dots,n, \end{equation*} while the conditions~$[\tilde{D}_i,\tilde{D}_j]=0$ will acquire the form \begin{equation*} D_i(X_j)-D_j(X_i)+[X_i,X_j]=0,\qquad 1\le i<j\le n. \end{equation*} In other words, we arrive to the classical definition of a ZCR (cf.~\cite{AblowitzKaupNewellSegur:InSTFAnNP}). \end{coordinates} \begin{example} \label{exmp:gjsis_nonloc:15} The well known two-dimensional ZCR for the KdV equation (see~\cite{AblowitzKaupNewellSegur:InSTFAnNP}) is given by \begin{equation*} \tilde{D}_x=D_x+A,\qquad\tilde{D}_t=D_t+B, \end{equation*} where \begin{equation*} A= \begin{pmatrix} 0&1\\ -u+\lambda&0 \end{pmatrix},\qquad B= \begin{pmatrix} -u_x&2u-4\lambda\\ -u_{xx}-2u^2+2\lambda u+4\lambda^2&u_x \end{pmatrix}. \end{equation*} \end{example} \begin{example}[vacuum Einstein equations] \label{exmp:gjsis_nonloc:16} Consider the Lewis metric~$\mathinner{\!}\mathrm{d} s^2=2f(x,y)\mathinner{\!}\mathrm{d} x\mathinner{\!}\mathrm{d} y+\sum_{j\le j}g_{ij}\mathinner{\!}\mathrm{d} z^i\mathinner{\!}\mathrm{d} z^j$ in~$\mathbb{R}^4$ with coordinates~$x$, $y$, $z^1$, and~$z^2$ (see~\cite{Lewis:SSSEqAxSGF}). The the vacuum Einstein equations read \begin{equation} \label{eq:gjsis_nonloc:27} (\sqrt{\det g}g_xg^{-1})_y+(\sqrt{\det g}g_yg^{-1})_x=0. \end{equation} After a re-parameterisation,~\eqref{eq:gjsis_nonloc:27} acquires the form \begin{equation} \label{eq:gjsis_nonloc:28} u_{xy}=\frac{u_xu_y-v_xv_y}{u}-\frac{1}{2}\frac{u_x+u_y}{x+y},\qquad v_{xy}=\frac{v_xu+y+u_xv_y}{u}-\frac{1}{2}\frac{v_x+v_y}{x+y}. \end{equation} B\"{a}cklund transformations and ZCR for~\eqref{eq:gjsis_nonloc:28} were constructed in many papers (see, e.g.,~\cite{BelinskiiZakharov:InEinEqMInSPTCExSS,Harrison:BTErEqGR,Maison:ArSAxSEinEqCIn,Maison:CInSAxSEinEq}). The latter is of the form \begin{equation*} \tilde{D}_x=D_x+A,\qquad\tilde{D}_y=D_y+B, \end{equation*} where \begin{equation} \label{eq:gjsis_nonloc:29} A=\frac{1}{2} \begin{pmatrix} -\frac{(\theta+1)u_x}{u}&\frac{(\theta+1)v_x}{u^2}\\ (\theta-1)v_x,&\frac{(\theta+1)u_x}{u} \end{pmatrix},\qquad B=\frac{1}{2\theta} \begin{pmatrix} -\frac{(\theta+1)u_y}{u}&\frac{(\theta+1)v_y}{u^2}\\ (1-\theta)v_y&\frac{(\theta+1)u_y}{u} \end{pmatrix}; \end{equation} here~$\theta=\sqrt{(\lambda+y)(\lambda-x)}$ and~$\lambda$ is the spectral parameter. Using ZCR~\eqref{eq:gjsis_nonloc:29}, a three-dimensional covering over~$\T(\mathcal{E})$ can be constructed (see~\cite{Marvan:ROpVEinEqS}). Let~$U$ and~$V$ be the variables in~$\T(\mathcal{E})$ corresponding to~$u$ and~$v$, respectively, and~$w^1$, $w^2$, $w^3$ be the nonlocal variables. Then the covering is given by the relations \begin{align*} w_x^1&=\frac{1-\theta}{2}v_xw^2+\frac{1+\theta}{2u^2}v_xw^3- \frac{1+\theta}{2u}U_x+\frac{1+\theta}{2u^2}u_xU,\\ w_x^2&=-\frac{1+\theta}{u^2}v_xw^1-\frac{1+\theta}{u}u_xw^2- \frac{1+\theta}{u^3}v_xU+\frac{1+\theta}{2u^2}V_x,\\ w_x^3&=(\theta-1)v_xw^1+\frac{1+\theta}{u}u_xw^3+\frac{\theta-1}{2}V_x \end{align*} and \begin{align*} w_y^1&=\frac{\theta-1}{2\theta}v_yw^2+\frac{1+\theta}{2\theta u^2}v_yw^3+ \frac{1+\theta}{2\theta u^2}u_yU-\frac{1+\theta}{2\theta u}U_y,\\ w_y^2&=-\frac{1+\theta}{\theta u^2}v_yw^1-\frac{1+\theta}{\theta u}u_yw^2- \frac{1+\theta}{\theta u^3}v_yU+\frac{1+\theta}{2\theta u^2}V_y,\\ w_y^3&=\frac{1-\theta}{\theta}v_yw^1+\frac{1+\theta}{\theta u}u_yw^3+ \frac{1-\theta}{2\theta}V_y. \end{align*} This covering gives rise to a B\"{a}cklund transformation of the form \begin{equation*} \theta U'=2uw^1+U,\qquad\theta V'=-u^2w^2-w^3, \end{equation*} i.e., to a recursion operator for symmetries. \end{example} \begin{remark} Note that with an arbitrary covering~$\tau\colon\tilde{\mathcal{E}}\to\mathcal{E}$ one can naturally associate a linear covering~$\tau^v\colon\T^v\tilde{\mathcal{E}}\to\tilde{\mathcal{E}}$. The space~$\T^v\tilde{\mathcal{E}}$ is a submanifold in~$\T\tilde{\mathcal{E}}$ and consists of tangent vectors that vanish under the action of the differential~$\tau_*$. Note also the existence of the exact sequence of coverings \begin{equation*} \xymatrix{ 0\ar[r]&\T^v\tilde{\mathcal{E}}\ar[r]\ar[dr]_{\tau^v}& \T\tilde{\mathcal{E}}\ar[d]^{\tilde{\tau}}\ar[r] &\T^s\tilde{\mathcal{E}}\ar[r]\ar[dl]^{\tau^s}&0\\ &&\tilde{\mathcal{E}}\rlap{,}&& } \end{equation*} where~$\tau^s\colon\T^s\tilde{\mathcal{E}}\to\tilde{\mathcal{E}}$ is the quotient. The terms of this sequence possess the following characteristic property: integrable sections\footnote{We say that a section is integrable if it preserves the Cartan distributions.} of~$\tau^v$ are infinitesimal gauge symmetries of~$\tau$, integrable sections of~$\tilde{\tau}$ are, as it was mentioned above, nonlocal $\tau$-symmetries of~$\mathcal{E}$, and integrable sections of~$\tau^s$ are $\tau$-shadows. \end{remark} \section*{Concluding remarks} \label{sec:gjsis_concl:concluding-remarks} We described a geometrical approach to partial differential equations which proved to be efficient both from the theoretical point of view and in a lot of applications. Based on this approach, in particular, Hamiltonian formalism for arbitrary normal ($2$-line) equations is constructed. On the other hand, a number of interesting and important problems are waiting for their solution. We intend to continue the research along the following lines: \begin{itemize} \item Generalisation of the Hamiltonian formalism from normal equations to arbitrary $p$-line ones that, in particular, include gauge-invariant systems. \item Incorporation of Dirac structures into the above described scheme and elaboration of their computation and use. \item Further development of the nonlocal theory and, in particular, analysis of differential coverings over the systems with the more than two independent variables and generalisation of the theory of variational brackets to nonlocal structures. \end{itemize} \section*{Acknowledgements} This work was supported in part by the NWO-RFBR grant 047.017.015, RFBR-Consortium E.I.N.S.T.E.IN grant 09-01-92438 and RFBR-CNRS grant 08-07-92496. We express gratitude to the participants of our seminar at the Independent University of Moscow (\url{http://gdeq.org}), where the topics of our paper were discussed. We are grateful to our colleagues whom we collaborated with for years and especially to Paul Kersten from Twente University (The Netherlands), Michal Marvan from the Silesian University in Opava (Czech Republic), Raffaele Vitolo from Salento University (Italy), and Sergey Igonin from the Utrecht University (The Netherlands). We are also grateful to Victor Kac for a number of useful remarks. We express our special thanks to an anonymous reader for his/her attention to the first version of this text and suggestions to improve the exposition. \providecommand{\MR}{\relax\ifhmode\unskip\space\fi MR } \providecommand{\MRhref}[2]{% \href{http://www.ams.org/mathscinet-getitem?mr=#1}{#2} } \providecommand{\href}[2]{#2} \providecommand{\urlprefix}{URL } \providecommand*{\eprint}[2][]{% \href{http://arXiv.org/abs/#2}{\begingroup \Url{arXiv:#2}}% }
\section{Introduction} A number of observations have established that the expansion of the universe is accelerating at late times ~\citep {1999ApJ...517..565P, 1998AJ....116.1009R, 1998ApJ...509...74G, 2003ApJ...598..102K, 2003ApJ...594....1T, 2004ApJ...607..665R, 2006A&A...447...31A, 2007ApJ...666..694W, 2009arXiv0901.4804H}. The cause of acceleration is usually attributed to an otherwise unobserved component called dark energy, but models of dark energy are generically plagued by fine-tuning issues ~\citep {1989RvMP...61....1W, 1992ARA&A..30..499C, 2000astro.ph..5265W, 2001LRR.....4....1C, 2004mmu..symp..235C}. One can also interpret these observations as a consequence of the gravitational dynamics being different from the evolution of a standard FRW universe under general relativity. Such differences could arise due to the symmetries of the FRW universe being broken in the real universe, and the assumptions of smallness of the perturbations being invalid ~\citep {2000PhRvD..62d3525B, 2005PhRvD..71b3524K, 2005PhLA..347...38E, 2006NJPh....8..322K}, or because General Relativity is not a correct description of gravity ~\citep {2000PhLB..485..208D, 2005PhRvD..71f3513C, 2008PhRvD..78f3503J}. With such fundamental questions at stake, a prime objective of physical cosmology is to understand the source and nature of this acceleration. All available current data ~\citep {2001ApJ...553...47F, 2005MNRAS.362..505C, 2006PhRvD..74l3507T, 2007MNRAS.381.1053P, 2008arXiv0803.0547K, Dunkley:2008ie, 2008AJ....135..512O, 2008ApJ...686..749K, 2009arXiv0901.4804H} is consistent with an FRW universe having dark energy in the form of a cosmological constant, yet various models of different classes are still allowed by the data. Therefore an important objective of current and future observational efforts is to study the acceleration of the universe in different ways and detect departures in the behavior from that expected in a standard $\Lambda$CDM model. In order to compute parameter constraints from observational data, one usually parametrizes our ignorance about dark energy with a time dependent equation of state (EoS) of dark energy as a specific function of redshift and theoretically computes the observational signatures. A very widely used choice, following the recommendations of the Dark Energy Task Force ~\citep{2006astro.ph..9591A}, is the CPL parametrization of the Equation of state ~\citep{2001IJMPD..10..213C,2003PhRvL..90i1301L}. This results in joint constraints on different parameters of the cosmological model, including the parameters of the EoS of dark energy. It is important to use different sets of observational data. Different kinds of data sets probe different physical imprints of dark energy leading to distinct shapes of constraints on parameters. Consequently, the simultaneous use of many `complementary' probes leads to the tightest constraints on cosmological parameters ~\citep{1998astro.ph..5117T,1998astro.ph..4168T,1999ApJ...518....2E,2003PhRvD..67h3505F}. Moreover, as indicated above, we can hardly be certain that the specific parametrization of the EoS chosen, or even the choice of the physical model causing the acceleration is correct. In that light, probing the observable effects of dark energy in terms of different physical aspects is even more important. A tension between constraints computed from different subsets of available data may be indicative of an incorrect parametrization ~\citep{2006astro.ph.11178C}, or even an untenable choice of a physical model. Traditionally, the main observations used to constrain cosmological parameters have pertained to the apparent magnitudes of Type IA supernovae, the power spectrum of anisotropies of the Cosmic Microwave Background (CMB), and the power spectrum of inhomogeneities in the matter distribution (matter power spectrum). The constraints from the supernovae relate to effects on the geometry of the universe due to dark energy through the changes in the background expansion. The CMB and matter power spectrum constraints stem mostly from a measurement of the geometry through the angular location of peaks of the anisotropy power spectrum and the peak positions of the Baryon Acoustic Oscillations (BAO), but also its effects on the growth of perturbations through the magnitude of the power spectrum. Current status of the parameter constraints on the basis of recent CMB, LSS, SNE observations can be found in~ \citep{2008PhRvD..77l3525W,2008PhRvD..78h3524X,2009arXiv0903.2532B}. Further, the use of observations of clusters of galaxies and weak lensing can be used to measure the growth of perturbations. It is therefore important to use probes of different aspects of cosmic evolution for constraining the cosmological parameters and models. From the viewpoint of both these perspectives, new probes for studying dark energy parameters are invaluable. In the above mentioned probes of the growth of cosmic structures, one studies the dependence of the dynamical growth of fluctuations on the cosmological parameters through the dependence of the growth of the amplitude (ie. size) of the fluctuations on the cosmology. However, in standard cosmology, while the fluctuations are stochastically isotropic, the individual fluctuations are not isotropic. Thus, a measure of the anisotropy and the time evolution of such measures can depend on cosmology in a distinct way. Consequently, this may be used to further constrain cosmological parameters. One expects that the signatures of anisotropic measures in observations would be related to the shapes of observed structures. Studying the evolution of shapes of high density regions (observable as galaxies or galaxy clusters at late times) and comparing with theory (eg.~\citep{2006ApJ...647....8H}) is difficult because this requires high resolution numerical simulations capturing the non-linear evolution of these systems. This difficulty can be avoided to a large extent by studying voids using semi-analytic methods. Therefore, the shapes of voids can be used to probe cosmology through the evolution of the anisotropy of fluctuations during cosmic growth. Park and Lee~\citep{2007PhRvL..98h1301P,2007arXiv0704.0881L} identified the probability distribution of a quantity which they called ellipticity \footnote{We note that this is not the conventional definition of ellipticity. Nevertheless, this is a convenient measure of the departure from spherical symmetry. Following ~\citet{2007PhRvL..98h1301P}, we shall refer to it as the ellipticity in the rest of the paper} related to the eigenvalues of the tidal tensor. They showed that the distribution was sensitive to the dark energy equation of state. Besides, they stated that the ellipticity could be derived from a catalog of galaxies, identifying voids of different sizes and measuring their shapes, and the distribution was verified using results from N-body simulations. This ellipticity is an example of a measure of anisotropy of individual fluctuations. The comparison of the probability distribution can provide complementary constraints on dark energy parameters if its cosmology dependence is different from other probes. We will not require new probes to study constraints from voids, rather one can study them using probes designed to study large scale structure in conventional ways, thereby allowing for better leveraging of data. Voids may be detected by the use of different void identification algorithms ~\citep{1997ApJ...491..421E,2001astro.ph.10449H,2002ApJ...566..641H,2008MNRAS.386.2101N,2008MNRAS.387..933C}, which find voids using different characteristics, and may be considered to be different definitions of voids. Properties of voids have been explored in 2dF~\citep{2004ApJ...607..751H} in SDSS~\citep{2005ApJ...621..643G,2007AstL...33..499T}. The shapes and sizes of voids in the SDSS DR5 have been explored in ~\citet{2009arXiv0904.4721F}. The main objective of this paper is twofold: (a) we want to quantify the potential of using void ellipticities to probe the nature of dark energy in terms of constraints on dark energy parameters, (b) and to clarify the model assumptions that are important for this procedure, which should be verified, or modified according to results from simulations. This paper is organized as follows: In Sec. II we review the idea that the shapes of voids can be quantified in terms of asymmetry parameters that can be related to the tidal tensor. We discuss the initial distribution of eigenvalues of the tidal tensor, and their evolution to study the evolution of the asymmetry parameters of voids and their dependence on the underlying cosmology. There are different theoretical choices of models to approximate the non-linear evolution of the initial potential field to observable void ellipticities. We discuss two different choices in the appendix and show that our results are insensitive to these choices. In Sec. III, we discuss the parameters from the surveys considered and our method of estimating the number of voids identified from these surveys. In Sec. IV, we write down a likelihood and explicit formulae for the Fisher matrix and use them to forecast constraints from these surveys. We also study how the constraints are degraded by systematic issues. We summarize the paper and discuss our outlook in Sec. V. \section{Theory} In this section, we outline the basic idea of using asymmetry parameters describing the shapes of voids in estimating cosmological parameters. The anisotropy of fluctuations may be captured by studying the eigenvectors and eigenvalues of the tidal tensor, which may be visualized as an ellipsoid with its principal axes along the eigenvectors of the tidal tensor, and sizes of the principal axes equal to the eigenvalues of the tidal tensor. At early times, the distribution of these eigenvalues at any point in space is known, and their evolution can be studied by semi-analytic methods. Therefore, the distribution of these quantities may be computed theoretically and it is desirable to find observational signatures of this distribution. Voids form around the minima in the density field of matter. The void geometry may be approximated by an ellipsoidal shape, which we shall refer to as the void ellipsoid. The central idea of ~\citet{2007PhRvL..98h1301P} is that the shape of the void ellipsoid as quantified by relative sizes of its principal axes is set by the geometry (functions of the eigenvalues) of the tidal ellipsoid and these should be strongly correlated. This implies that the ellipticity measured from the geometry of voids can be used as an observable for specific functions of the eigenvalues of the tidal tensor. Observations of void shapes at different redshifts can then be used to trace the evolution of the stochastic distribution of these eigenvalues of the tidal ellipsoid at different redshifts. This contains dynamical information that may be used to constrain cosmological parameters. We briefly describe measures of ellipticity of the void ellipsoid and their connection to the eigenvalues of the tidal ellipsoid in subsection ~\ref{ansatz}: this specifies the functions of the tidal eigenvalues that are constrained by the void shapes. We then describe the distribution of eigenvalues of the initial tidal tensor appropriate to an observed void in subsection ~\ref{distribution_eigenvalues}. Then, in appendices ~\ref{appendix:gesf} and ~\ref{appendix:EllipsoidalCollapse}, we study the time evolution of the initial eigenvalues using two different approximations, and find them to be consistent. \subsection{Relating the Asymmetry Parameters to the tidal tensor} \label{ansatz} To describe the dynamics, we choose the comoving coordinates of particles (or galaxies) as the Eulerian coordinates $\vec{x}$, while the Lagrangian coordinates are taken to be $\vec{q}$, which are approximately the `initial' Eulerian coordinates at some chosen large redshift. The two coordinates are always related through the displacement field $\vec{\Psi}(\vec{q},\tau)$. \begin{equation} \vec{x} =\vec{q} +\vec{\Psi}(\vec{q},\tau) \label{LagrangianMapping} \end{equation} While the solution $\Psi(\vec{q},\tau)$ describes the dynamics completely, partial aspects of the dynamics may be described by other measures. The asymmetry of the fluctuation can be understood in terms of the eigenvectors and eigenvalues of the tidal tensor $T_{i,j} =\frac{\partial \Psi_i(\vec{q})}{\partial q_j}.$ This can be visualized as an ellipsoid, which we shall refer to as the tidal ellipsoid, with principal axes along the eigenvectors of the tidal tensor with sizes equal to the eigenvalues. For a spherically symmetric fluctuation, these eigenvalues are equal, while the departure from spherical symmetry may be characterized by different choices of functions of ordered eigenvalues of the tidal tensor. (See Appendix~\ref{OtherParametrizations} for some other popular choices in the literature.) This was recognized and used in correcting for ellipsoidal collapse of halos rather than spherical collapse in Press-Schechter like estimates of the mass function of dark matter halos ~\citep{2001MNRAS.323....1S,2002MNRAS.329...61S,2001ApJ...555...83C}. From a theoretical side, we can describe the evolution of the distribution of these eigenvalues. Therefore, it is these dynamical quantities that we are interested in, even though they are not directly observable. We will next proceed to describe observable quantities which relate to the shape of the voids, and then show how functions of those observables trace functions of these dynamical quantities. Since voids form around minima of the density fields where the gradient of field vanishes, one can approximate the density profiles around the minima by truncating the Taylor expansion at second order. This gives density profiles that are ellipsoidal in shape. One may expect voids to inherit this shape, and therefore be approximately ellipsoidal. In fact, voids have often been modeled as spherical (eg.~\citep{2004ogci.conf...58V}), while others have argued that the shapes of larger voids fit ellipsoids well only for smaller voids~\citep{2006MNRAS.367.1629S}. For irregularly shaped voids (obtained by suitable void identification algorithms), one can define a void ellipsoid by fitting a moment of inertia tensor to the positions of observed void galaxies $\vec{x}$ in Eulerian coordinates relative to the void center $\vec{x}^v$ $$S_{ij} = \frac{\sum_k (x^k_i - x^v_i)(x^k_j-x^v_j)}{\mathcal{N}},$$ where the index k runs over the observed galaxies in the void region, and $\mathcal{N}$ is the number of galaxies fitted. The void ellipsoid can be defined as the ellipsoid with principal axes along the eigenvectors of this mass tensor, and lengths proportional to the square root of the eigenvalues $\{J_1,J_2,J_3\}$. Here, we shall ignore the discrepancy between the actual shape and this void ellipsoid. Following ~\citet{2007PhRvL..98h1301P} (see Appendix.~C of \citet{2009arXiv0906.4101L} for a calculation to first order), one can relate the eigenvalues of the tidal tensor $\{\lambda_1,\lambda_2,\lambda_3\}$ to the functions of the ratio of eigenvalues of the void ellipsoid which were called ellipticity. Accordingly, the ellipticities $\{\epsilon,\omega\}$ of the void ellipsoid are to first order \begin{equation} \label{eqn:VoidEllipsoidEllipticity} \epsilon = 1 - \left(\frac{J_1}{J_3}\right)^{1/4} \approx 1 -\left(\frac{1- \lambda_1}{1-\lambda_3}\right)^{1/2}, \qquad \omega = 1 - \left(\frac{J_2}{J_3}\right)^{1/4} \approx 1 -\left(\frac{1- \lambda_2}{1-\lambda_3}\right)^{1/2}. \end{equation} Clearly, this relation will be affected, at least to some extent, by more detailed dynamics. This would lead to $\epsilon$ measured from data sets on voids being correlated with the functions of $\{\lambda_i\}$ with some scatter. In computing parameter constraints, we shall account for this in terms of a variance in the quantity $\epsilon$ which also contains contributions from observational errors. We shall assess the impact of this assumption of the void shapes being perfect tracers of the eigenvalues by studying the degradation of constraints on increasing the variance in our study of systematics in Section.~\ref{ResultsIssues}.\\ \subsection{Distribution of Initial Eigenvalues of the Tidal Tensor} \label{distribution_eigenvalues} An observed void evolves from a fluctuation of low underdensity at early times when the distribution of fluctuations was Gaussian. Given a void of a given density contrast, at a particular redshift, we wish to calculate the distribution of eigenvalues of the tidal tensor of the initial fluctuation. At early times, the fluctuations are small enough, their growth can be described by linear perturbation theory, and the distribution remains Gaussian. One can use the statistical properties of filtered isotropic and homogeneous Gaussian fields to derive a probability distribution of the ordered eigenvalues of the tidal tensor given by the Doroshkevich formula. \begin{equation} \label{DKV} P(\lambda_1, \lambda_2, \lambda_3\vert \sigma_R ) = \frac{3375}{8\sqrt{5}\sigma_{R}^6} \exp\left(-\frac{-3K_1^2}{2\sigma^2_{R}} + \frac{15 K_2}{2\sigma^2_{R}}\right) K_3 \end{equation} where $K_1 =\lambda_1 +\lambda_2+\lambda_3,\quad K_2 =\lambda_1\lambda_2 +\lambda_2\lambda_3+\lambda_3\lambda_1$, while $K_3 =-(\lambda_1-\lambda_2)(\lambda_2-\lambda_3)(\lambda_3-\lambda_1),$ and $\sigma^2_R$ is the variance of the smoothed overdensity field at the filtering scale $R$ at that time. Note, that this gives the distribution of the size of the eigenvalues over all spatial points. This distribution is extremely similar but slightly different if restricted to the maxima of the Gaussian field ~\citep{1986ApJ...304...15B}, or the minima of the Gaussian field ~\citep{2009arXiv0906.4101L} which should evolve to voids. For the small fluctuations, one can use the Jacobian of the transformation from Eulerian to Lagrangian coordinates to show that the sum of the eigenvalues $K_1$ can be identified with the density contrast.\\ It should be noted that this distribution depends on the filtering scale $R_{\text{Smooth}}$ as a parameter while the size of voids is not important. This is appropriate for comparison with a dataset of voids obtained from redshift surveys by means of an algorithm which uses a filtering scale as a parameter, rather than the void size. This is true for a class of algorithms that define voids as regions of space where the smoothed matter density is a minimum (eg.\citep{2007MNRAS.375..489H,2009arXiv0906.4101L}) with the smoothing scale $R_{ \text{Smooth}}$ being a parameter, with the actual size of voids not being crucial to the definition. On the other hand there are Void Finding algorithms which define voids as the largest contiguous underdense regions, obtained by some form of clustering algorithms. A corresponding parameter here is the size $R$ of the voids related to the void volume by $R^3\equiv \frac{3 V}{4\pi},$ while the smoothing scale is not crucial. While each algorithm might yield slightly different properties of voids, it would be expected that they are not too different. In Appendix~\ref{appendix:gesf}, we show that a calculation based on the generalized excursion set formalism can be used to calculate the distribution of eigenvalues of an initial fluctuation that evolves to form a void of size $R$. The result of this calculation supports the above result. \subsection{Evolution of the Tidal Eigenvalues} \label{evolution_eigenvalues} At low redshifts, gravitational collapse introduces non-linearities into the evolution leading to non-Gaussian distributions of the density field. Thus, the distribution of the tidal eigenvalues of the previous subsection which assumed Gaussianity are not directly applicable. We study the evolution of these eigenvalues with time in two different methods, one based on the Zeldovich approximation and one based on \citet{1996ApJS..103....1B}. \\ It is well known that non-linearity is manifested much less in the displacement field or the gravitational (and the related displacement) potential than in the density field. Therefore before shell-crossing, the evolution of structures from initial condition may be described by the Zeldovich approximation, where the displacement field is assumed to be separable into a time dependent and time independent part. $\Psi(q,\tau) = {D(\tau)\over D(\tau_0} \Psi(q,\tau_0)$, where $D(\tau)$ is the linear growth function. Hence, at a particular spatial point, its eigenvalues $\lambda_i(\tau)$ at time $\tau$ evolve linearly from the eigenvalues $\lambda_i(\tau_0)$ at some initial time $\tau_0$ as $\lambda_i (\tau)= D(\tau)\lambda_i(\tau_0)/D(\tau_0).$ Rewriting the early time eigenvalues in the Doroshkevich formula (Eqn.~\ref{DKV}) in terms of the eigenvalues at time $\tau,$ one can then find a distribution of eigenvalues at any time to be given by the Doroshkevich formula where the $\sigma_R$ is replaced by $D(\tau)\sigma_{R}/D(\tau_0)$, the linearly extrapolated variance over the Lagrangian smoothing scale $R$. The formula is exactly the same as Eqn.~\ref{DKV} with the variance $\sigma^2_R$ being replaced by the linearly extrapolated variance $\sigma^2(R,z)$, and $\lambda_i$ replaced by the eigenvalues at the redshift of the void. Further, since the sum of the eigenvalues $K_1$ at early times was equal to the density contrast at that time, the term $K_1$ is equal to the linearized density contrast of the time of the void \begin{equation} \delta_{lin}(\tau) = \frac{D(\tau)}{D(\tau_0)}\delta(\tau_0) =\frac{D(\tau)}{D(\tau_0)} (\lambda_1(\tau_0)+\lambda_2(\tau_0)+\lambda_3(\tau_0)) =(\lambda_1(\tau)+\lambda_2(\tau)+\lambda_3(\tau)) \label{eqn:deltalin} \end{equation} In regions of high density peaks where structure forms, it has been found that modeling the density growth as a collapse of a homogeneous ellipsoid leads to a better approximation to N body simulations. It is unclear whether this should also be true for low density regions like voids. In Appendix~\ref{appendix:EllipsoidalCollapse}, we study the evolution of the eigenvalues of the tidal tensor based on ellipsoidal collapse ~\citep{1996ApJS..103....1B} and find the differences with the evolution computed using Zeldovich approximation to be small. \subsection{Cosmology Dependence of the Distribution of Ellipticity} Therefore, using the Zeldovich approximation, one can write down the probability distribution of the eigenvalues of the tidal tensor at any time. Further, using the relations of the ellipticities of the void (Eqn. ~\ref{eqn:VoidEllipsoidEllipticity}) and the relation of the linearly extrapolated density contrast to the eigenvalues $\{\lambda_1,\lambda_2,\lambda_3\}$, one can recast this as the joint distribution of the ellipticities $\{\epsilon,\omega\}$ given the smoothing scale and the linearly extrapolated density contrast. Following Park and Lee, we define $\mu,\nu$ and write the probability distribution for the larger ellipticity $\epsilon$ \begin{eqnarray} \label{eqn:ellipticitydist} \mu & = &\left(J_2/J_3\right)^{1/4}, \qquad \qquad \nu = \left(J_1/J_3\right)^{1/4} \nonumber\\ P(\mu, \nu \vert \sigma_{lin} (R,z), \delta_{lin}(z)) &=& \frac{3^4/4}{\Gamma(5/2)}\left(\frac{5}{2~\sigma^2_{lin}(R,z)}\right)^{5/2} \exp \left( -\frac{5 \delta_{lin}^2(z)}{2~\sigma_{lin}^2(R,z)} + \frac{15 K^{\delta}_{2}}{2~\sigma_{lin}^2(R,z)} \right) K^{\delta}_3 J\nonumber\\ P(\epsilon \vert \sigma_{lin}(R,z),\delta_{lin}(z)) & = & \int_{1-\epsilon}^1 d\mu P(\mu,1-\epsilon \vert \sigma_{lin} (R,z), \delta_{lin}(z)) \end{eqnarray} where $K^{\delta}_{2},K^{\delta}_{3}$ are the values of $K_2,K_3$ in Eqn.~\ref{DKV} in terms of $\mu,\nu$ when the constraint of Eqn.~\ref{eqn:deltalin} holds, and $J$ is the Jacobian in the transformation from the coordinates $\{\lambda_1,\lambda_2,\delta_{lin}\}$ to $\{\mu,\nu,\delta_{lin}\}.$ This last equation gives the probability distribution of the larger ellipticity $\epsilon$ marginalized over the smaller ellipticity $\omega$. It depends on the cosmology only through the linearly extrapolated variance $\sigma^2_{lin}(R,z)$ of density fluctuations $\delta(x,z)$ smoothed at a certain filtering scale $R$ by a window function $W_R(x,x^\prime).$ \begin{equation} \sigma^2_{lin}(R,z)\equiv \lept{\delta^\star_{R}(x,z)\delta_{R}(x,z)} =D^2(\tau)\sigma^2_{R} \qquad \delta_{R}(x,z) =\int d^3x^\prime \delta (x,z) W_{R}(x,x^\prime) \end{equation} where $D(\tau)$ is the growth function and $\sigma_R$ is evaluated at early times. For qualitative understanding, it is useful to think of the variance depending on cosmology through $\sigma_R$ which depends on the primordial power spectrum and the wave mode dependent transfer function, and the subsequent scale independent growth described by the growth function $D(\tau)$. While the transfer function depends on most of the cosmological parameters, in most models dark energy does not become significant at early times. Therefore most of the effects of dark energy are embedded in the growth function. Closed analytic forms for the growth function are not known for non-flat cosmologies, with time varying equations of state dark energies, but~\citet{2005A&A...443..819P} improves upon a fit to the growth function by~\citet{2003ApJ...590..636B}, so that the fit works for non-flat cosmologies having dark energy with time varying equation of state as long as they are close to flat LCDM models, even when the equation of state is less than -1. If we consider the CPL parametrization \begin{equation} w(z) = w_0 + w_a \frac{z}{z+1},~ \end{equation} we see in the left panel of Fig.~\ref{GrowthDerivs} that the growth function changes more dramatically as a function of $w_0$ than for $w_a$ for redshifts below unity. Thus, we expect, that constraints from voids in these redshift ranges should be stronger on $w_0$ than on $w_a$. From the right panel of Fig.~\ref{GrowthDerivs}, we can see the effect of the filtering scale $R$ on the distribution. Since $\sigma_{lin}(R,z)$ is a monotonically decreasing function of $R,$ a larger filtering scale (a) shifts the distribution towards smaller values of $\epsilon,$ and (b) sharpens the distribution. This is consistent with intuition based on previous studies ~\citep{1979ApJ...231....1W,1984MNRAS.206P...1I,1996MNRAS.281...84V}. Leaving all other variables the same, increasing $R$ corresponds to excluding the smaller voids. Since the variation of possible values is caused by the variance in the Gaussian distribution, a smaller value of $\sigma_{lin}(R,z)$ also corresponds to a sharper distribution.\\ In this paper, we shall assume that all voids are found at a linearized density contrast of $\delta_{lin} =-2.81,$ the underdensity at shell crossing. We shall compute $\sigma_{lin}(R,z)$ directly from numerical integration of the smoothed density fluctuations evolved by a modified version of the Boltzmann code \verb CAMB ~\citep{2000ApJ...538..473L}. \begin{figure*}[!htp] \begin{center} \includegraphics[width=0.9\textwidth]{Fig1} \caption{Left Panel:The Derivative of the growth function with respect to the dark energy parameters $w_0$, and $w_a$. The growth function shown has been normalized to unity at a redshift of 0.01. Right Panel: The theoretical distribution of the largest ellipticity $\epsilon$ as a function of $\sigma(R,z)$ for $\delta_{lin}$=-2.81} \label{GrowthDerivs} \end{center} \end{figure*} \section{Distribution of Ellipticity: Connecting to Observations} \subsection{Estimate of Voids to be found from a survey} Next, we proceed to estimate the number of voids that we expect to find in a certain survey. We model a survey by considering a redshift survey, which can measure the redshifts of the galaxies up to a limiting visual magnitude of $m_L$ in a given filter and from a minimum redshift of $z_{min}$ to a maximum of $z_{max}$. In case of photometric surveys, the errors in redshift can be much larger, leading to errors in the size of the ellipse along the line of sight, consequently the distribution of ellipticities will have to be marginalized over this error. Here, we will limit our considerations to spectroscopic surveys, where the error in measuring the redshift of the galaxies $\sim 10^{-4}$ is negligible. In order to estimate the number of voids of a particular size at a particular redshift, we use the Press-Schechter formalism to determine the number density of voids in a redshift bin centered at $z,$ with Eulerian comoving radius between $R_E$ and $R_E + d R_E.$ Simulations indicate that the number density of voids peaks at a density contrast of $\delta \approx -0.85$~\citep{2007PhRvL..98h1301P}, we shall consider all the voids to have a density contrast of 0.8, which can be seen to correspond to a linearly extrapolated density contrast of -2.81 using the fitting function in ~\citet{1996MNRAS.282..347M}. While the usual Press-Schechter formalism matches simulations well at redshift ranges below $\approx 2,$ it fails to predict the number of voids correctly at small scales due to the `void in cloud problem' , which can be avoided if at each redshift, we restrict ourselves to scales larger than the non-linearity length scale (Lagrangian) $R_{min}^{VinC}(z)$ where $\sigma (R_{min}^{VinC}(z),z)= 1$~\citep{2004MNRAS.350..517S}. Then, the Press-Schechter formalism reliably predicts the number of voids with the replacement $\delta_c = 1.69 \rightarrow \delta_v =-2.81$ in the standard Press-Schechter formalism ~\citep{1974ApJ...187..425P}. The number of voids of a particular size can then be found by integrating over the cosmological volume in the redshift bin, and over the range of radii allowed. \begin{eqnarray} n_v(R_E,z) dR_E~&=&~ \frac{3 }{2\pi R_E^3} P\left(-\frac{\vert\delta_v(z)\vert}{\sigma_{R_E}}\right) \left \vert \frac{d}{dR_E}\frac{-\vert\delta_v (z)\vert}{\sigma_{R_E}} \right \vert dR_E \\ N_{void}~&=&~ \int_z^{z+\Delta z} d\Omega dz \int_{R_E}^{R_E+\Delta R_E} d R_E \; \frac{dV}{dz d\Omega} \; n_v(R_E)\nonumber \end{eqnarray} where $P(y) = \sqrt{\frac{1}{2}}exp(-y^2/2).$ The number density of voids thus depends exponentially on $\sigma_R$ and therefore the number of voids is extremely sensitive to the minimum radius used. Since voids are detected by observing galaxies rather than the matter density, the number of voids detected with small radii will be strongly affected by shot noise (discussed in subsection \ref{EffectsOfShotNoise}). We therefore only consider voids with radii greater than a critical radius $R \geq R_{\mbox{min}}^{\mbox{shot}}(\mbox{z,Survey})$. For our purposes then, the minimum of the range of radii of voids at a redshift $z$ considered must be set to the maximum of $R_{min}^{Vinc}(z)$ and $R_{min}^{shot}(z,Survey).$\\ We now explain our method for computing $R_{min}^{shot}(z,Survey),$ from the parameters for a survey. The minimum radius of voids that we will consider should be related to the average separation of galaxies \emph{observed} $l_{\mbox{sep}}(z)$ at the redshift $z$ by the survey in question. We choose this relationship to be linear $R_{\mbox{min}}^{\mbox{shot}}(\mbox{z,Survey})= A l_{\mbox{sep}}(z),$ and relate the average separation to the average number density of observed galaxies $n^{bg}_{gal}(z)$ at that redshift for the survey. A choice of $A=2$ implies that the probability that a detected void is just due to shot noise is less than 0.5 percent while such a scenario for $A=1$ is of the order of 50 percent, though void identification algorithms can do better, since they can exploit the contrast between voids and their higher density environments. In any case, the interesting regime is in between these numbers and we shall later explore the sensitivity of constraints to this range. This background number density of observed galaxies $n^{bg}_{gal}(z)$ can be related to the survey parameters. The mean number density of galaxies in the background universe can be calculated from the luminosity function ~\citep{2001AJ....121.2358B} of galaxies at the filter band used in the survey by, \begin{equation} \label{nbggal} n^{bg}_{gal}(z)=\int_{-\infty}^{M_L} dM \Phi_X(M,z) \end{equation} where $\Phi_X$ is the luminosity function for the filter $X$ and $M_L$ is the limiting absolute magnitude of objects at redshift $z$ which are observed by the survey. It can be calculated from the limiting apparent magnitude of the survey $m_L$ by using the formula, \begin{equation} M_L= m_L -5 \log_{10}D_L(z) +5 - A(z) - K(z) \end{equation} Here $D_L(z)$ is the luminosity distance to the redshift $z$ in units of pc, $A(z)$ is the correction due to extinction and $K(z)$ is the K correction arising from the difference in the observed luminosity of and the rest frame luminosity of an object in a particular frequency band due to redshifting of photons. We note that $R_{min}^{VinC}$ depends on the cosmology, but is independent of the survey, while $R_{min}^{shot}(z,survey)$ also depends on the survey through the filter band, and the limiting magnitude. A plot of $R_{min}^{noise}$ and $R_{min}^{VinC}$ for surveys considered in this paper is shown in Fig.~\ref{Fig:PlotRmin}. Thus, our estimate of the number of voids identified by each survey depends on the cosmology, the value of the proportionality constant $A$ and the survey parameters. \begin{figure*}[!hpt] \begin{center} \includegraphics[width=0.45\textwidth]{Fig2a} \includegraphics[width=0.45\textwidth]{Fig2b} \caption{Setting the minimum size of voids: the dashed red curve shows the $R_{min}^{VinC}$, while the solid thin (thick) curves show the (twice) the average separation of observed galaxies for a SDSS DR7 like survey (left) and a EUCLID like survey (right). At a particular redshift, we only consider voids with sizes larger than both these scales.} \label{Fig:PlotRmin} \end{center} \end{figure*} \begin{table*} \label{VoidSurveys} \caption{Surveys and parameters used for estimating the number of voids that can be found by the survey. We chose a survey like SDSS DR7 as an example of a current survey, and EUCLID as an example of a futuristic survey. For reference, we show the number of galaxies that these surveys are expected to observe.} \begin{tabular}{|c|c|c|c|c|c|} \hline Survey & $f_{sky}$ & Freq Band & Limiting Magnitude & Number of Voids & Number of Galaxies\\ & & & & $A=2$,$A=1$ & \\ \hline \verb=SDSS= DR7\footnote{\url{http://www.sdss.org/dr7/coverage/index.html}} & 0.24 & r & 18 & 1292,3104 & 1.7 $10^6$\\ \hline \verb=EUCLID= \footnote{\url{http://hetdex.org/other_projects/euclid.php}} & 0.48 & K & 22 & 1.4 $10^5$, 2.3 $10^6$& 5.2 $10^8$ \\ \hline \end{tabular} \end{table*} \section{Results} \subsection{Likelihood function and Fisher matrix} In order to study the potential constraints on cosmological parameters, we need to write down a simple model for the data. We assume that by applying appropriate simulation algorithms, we can identify a set of voids at each redshift bin corresponding to a particular smoothing scale. We expect to measure the ellipticities of each of these voids with some error. We model the error as an additive Gaussian noise $n$ on the ellipticity $\epsilon_s$: \begin{equation} \epsilon_d (R,z) = \epsilon_s(R,z) + n, \quad n \sim G(0,\sigma_\epsilon ) \end{equation} $\epsilon_s$ itself is a random variable following the distribution of the ellipticities at the relevant redshift. Then we can write down the likelihood function, which is the probability for finding a void with a measured largest ellipticity $\epsilon_d$ given the cosmological parameters \begin{equation} L(\epsilon_d \vert \Theta ) = \int d\epsilon_s P(\epsilon_d \vert \epsilon_s) P(\epsilon_s \vert \sigma_\epsilon ,\Theta) \label{LikeIndividual} \end{equation} One expects that the error in measuring the ellipticities will be set by the errors in measuring the principal axes of the void ellipsoid. For a spectroscopic survey, the positions of galaxies are well measured. Ignoring effects of redshift distortion/finger of god effects the precision level of the measurement of the principal axes would be set by the errors in the void finding algorithm. Of course, this will be limited by the relative sizes of the void wall thickness to the void radius $\Delta$. For $\Delta \sim0.1-0.4$, $\epsilon \approx 0.2$ around the maximum for standard cosmological parameters, the error in $\epsilon$ is of the order of 0.1. The errors in the measurement of each void is statistically independent. Thus the likelihood function for an entire data set consisting of voids at different redshifts can be computed as the product of Eqn. \ref{LikeIndividual} for each void. Consequently, the log of the likelihood function ${\cal{L}}(\epsilon_d \vert \Theta )$ is additive for each void. Given the likelihood function for a single void, one can compute the Fisher matrix F defined as an expectation over all possible sets of data, \begin{equation} F_{ij}= \left\langle \frac{\partial{\cal{L}}(\epsilon_d \vert \Theta, \sigma_\epsilon)} {\partial \Theta_i} \frac{\partial{\cal{L}}(\epsilon_d \vert \Theta, \sigma_\epsilon)} {\partial \Theta_j} \right\rangle = \int_0^1 d\epsilon_d L(\epsilon_d \vert \Theta, \sigma\epsilon) \frac{\partial{\cal{L}}(\epsilon_d \vert \Theta, \sigma_\epsilon)} {\partial \Theta_i} \frac{\partial{\cal{L}}(\epsilon_d \vert \Theta, \sigma_\epsilon)} {\partial \Theta_j} \end{equation} where all the derivatives are taken at a fiducial choice of the cosmological parameters $\Theta_p$. Since, in our model the error in measuring the ellipticity is independent of the cosmological parameters, and the ellipticity depends on the cosmological parameters through the variance of the fluctuations $\sigma^2_R$ only, we can factorize this into a matrix of mixed partial derivatives of $\sigma_R$ with respect to the cosmological parameters, and the derivatives of the log likelihood with respect to $\sigma_R$. We evaluate both of these derivatives numerically. The main contribution to the derivatives comes from the regions where the probability is smallest. However, these contributions are suppressed in the expectation values, since these regions have low probabilities. Finally, we must sum this contribution for the Fisher matrix over all the voids in the data set. The result thus depends critically on the number of voids in the data set. \subsection{Forecasts of constraints on the CPL parameters} We consider Fisher forecasts for a cosmology with the non-baryonic matter assumed to be cold, neglect effects of neutrino masses and parametrize the evolution of the dark energy equation of state with a CPL parametrization. The primordial perturbations are assumed to be Gaussian distributed, and characterized by a spectrum which is a power law with an initial amplitude $A_s$, and a scale independent tilt $n_s$. The distribution of ellipticities depends on both the amplitude of primordial perturbations, and the spectral index through the dependence of the variance on the scale of smoothing. As is well known, these quantities $A_s, n_s$ are not exactly known, and have a degeneracy with $\tau,$ the optical depth of reionization. Further, the constraints on the equation of state parameters can depend strongly on the knowledge of the curvature parameter ~\citep{2009arXiv0903.2532B}. We therefore consider forecasts for constraints on the CPL parameters $w_0,w_a$ after marginalizing over all other cosmological parameters from a maximal set shown in Table.~\ref{Params}, along with the fiducial values used for computing the Fisher forecasts. All of these parameters are not well constrained by a single experiment. Consequently, we shall consider Fisher forecasts using ellipticity distribution of voids from two spectroscopic surveys: the recent \verb SDSS DR7 and the futuristic~\verb EUCLID with the survey parameters assumed summarized in Table.~\ref{VoidSurveys}. We will assume $A =1,$ $\sigma_\epsilon=0.1 .$ Following the work in ~\citep{2009arXiv0906.4101L}, we will identify the smoothing scale as being a quarter of the radius of the void. For CMB constraints, we will consider Fisher forecasts computed from \verb PLANCK ~ \footnote{\url{http://www.rssd.esa.int/index.php?project=Planck}} The expressions for the Fisher matrix for CMB data are given in ~\citet{1998astro.ph..4168T}. The survey parameters for \verb PLANCK ~are taken from the Table. 1.1 of the PLANCK~Bluebook~\citep{2006astro.ph..4069T}, and are summarized in Table.~\ref{PlanckParams}. We consider Fisher forecasts of Supernovae from two surveys: for a survey like Dark Energy Survey the number of supernovae expected is of the order of 1300, and the maximum redshift is around 0.7. We model this with a redshift distribution taken from~\citep{2008ApJ...675L...1Z} designed to be cut off at z=0.7, and assume perfect measurement of redshift, due to plans of spectroscopic follow-up. The errors in the magnitude are assumed to be of the order of the intrinsic dispersion from light curve fitting techniques today (0.15). We also consider a futuristic photometric Supernova IA survey \verb LSST ~\citep{2009arXiv0912.0201L}, where about 500,000 SNe IA suitable for constraining dark energy parameters could be observed. We model the errors by assuming magnitude errors of the order of 0.12 from intrinsic dispersion, and photometric errors in redshift determination of the order of $\Delta z=0.01(1+z),$ and assuming that this adds an error $\frac{dm}{dz} \Delta z$ in quadrature to the intrinsic dispersion. We use the redshift distribution in Table 1.2 of the~\citep{2009arXiv0912.0201L} to model the redshift distribution of the LSST survey. \begin{table*} \label{Params} \caption{Parametrization of the cosmology and the fiducial values chosen for the maximal set of parameters used in evaluating the Fisher forecasts. Constraints are also discussed after imposing flatness.} \begin{center} \begin{tabular}{|c|c|c|c|c|c|c|c|c|} \hline $\Omega_{\rm{b}}\rm{h}^2$ & $\Omega_{\rm{c}}\rm{h}^2$ & $\theta$ & $\tau$ & $\Omega_{\rm{k}}$ & $w_0$ &$w_a$ & $n_s$ & $log(10^{10}A_s)$ \\ \hline 0.02236 & 0.105 & 1.04 & 0.09 & 0.0 & -1 & 0 & 0.95 & 3.13 \\ \hline \end{tabular} \end{center} \label{PriorsTable} \end{table*} \begin{table*} \caption{Parameters of the PLANCK~Survey used in determining CMB constraints} \label{PlanckParams} \begin{tabular}{|c|c|c|c|c|c|c|c|} \hline Frequency Channel & 30 & 44 & 70 & 100 & 143 & 217 & 353\\ (GHz) & & & & & & & \\ \hline Beam Width & 33 & 24.0& 14.0 & 10.0 & 7.1& 5.0 & 5.0 \\ (FWHM) arc min & & & & & & & \\ \hline Temperature Noise per Pixel &2.0 & 2.7 & 4.7 & 2.5 & 2.2 & 4.8 & 14.7 \\ \hline Polarization Noise per Pixel & 2.8 & 3.9 & 6.7 & 4.0 & 4.2 & 9.8 & 29.8 \\ \hline \end{tabular} \end{table*} \begin{figure*}[!hpt] \begin{center} \includegraphics[width=0.95\textwidth]{Fig3} \caption{~{\textbf{Comparison of forecasts on one $\sigma$ constraints on the CPL parameters with standard probes} using the identification $R_{ \text{Smooth}}=R_{Void}/4$, $A=1$, and $\sigma_\epsilon=0.1$:} for data from the near future (left panels) and futuristic data (right panels). PLANCK and HST priors were used in all of these forecasts. For reference, we show the current constraints ~\citep{2009arXiv0903.2532B} in the thick green contours, and forecasted constraints from clusters (number counts and power spectrum) + PLANCK taken from ~\citep{2004PhRvD..70l3008W}. } \label{Constraints} \end{center} \end{figure*} In Fig.~\ref{Constraints}, we present the constraints on the equation of state parameters $w_0, w_a$ by combining constraints for two sets of data (a) data representative of current or near future (left panels), and (b) data representative of more futuristic data (right panels). The forecasts for one sigma constraints using void ellipticities + CMB + HST are shown in open circles, assuming $A=1.$ The error in measuring the ellipticities $\sigma_{\epsilon}$ is taken as 0.1. The ellipses made of black "+" show the constraints for SNe + HST +CMB (PLANCK). The solid, thick, blue ellipses show the constraints when these constraint are combined (CMB (PLANCK) +SNE +HST + Voids). In the left panels of the figure, the voids considered are from a survey like SDSS DR7, and the SNe considered are from a survey like DES. In the right panels the voids considered are from a futuristic survey like EUCLID, and the SNe are from a futuristic photometric survey like LSST. The upper panels show the constraints marginalized over all other parameters in the maximal set, while the lower panels show the marginalized constraints for a flat universe. For reference, we show the thick, green contours showing the one sigma constraints from current SNe (Union) + HST + CMB (WMAP 5) data from~\citep{2009arXiv0903.2532B}. For the flat universe in the lower panels , we also show the constraints from CMB (PLANCK) + HST + Clusters (Power spectrum + Number counts) from ~\citep{2004PhRvD..70l3008W}. In the lower left panel the Clusters considered from the SPT survey, while the lower right panels show the constraints from clusters from LSST. Firstly, these figures show that the inclusion of constraints from void ellipticities significantly improves parameter constraints and the constraints from Voids along with CMB and HST data are comparable to the joint constraints obtained by using Supernovae IA, CMB and HST data both in the near future and the far future. As is common, following ~\citet{2006astro.ph..9591A}, we quantify this in terms of a Figure of Merit (FoM) which is inversely proportional to the area of the two sigma contours (ie. proportional to the inverse of the determinant of the $w_0,w_a$ submatrix of the inverse of the Fisher Matrix). We calculate the FoM relative to the FoM without voids for each of the upper panels: $$FoM(\rm{experiments}) = det(\rm{SNE+PLANCK+HST})/det(\rm{experiments})$$ where $\rm{experiments}$ refer to the combination of experiments we consider the FoM for, and the SNE experiments in the numerator refer to the DES for the left panel, and LSST for the right panel. The relative FoM for these results are shown in ~\ref{table:FoM} (for $A=1,\sigma = 0.1$). We see that the constraints with the use of (Voids + CMB + HST) is not good as, but somewhat comparable (Relative FoM =0.6) to the constraints due to (SNe + CMB + HST), but adding the void constraints to the SNE +CMB +HST data offers a moderate gain (FoM = 13.3). For the futuristic case, the use of (Voids + CMB + HST) is is better than the corresponding (CMB + SNe +HST) data (FoM=70.4), while combining these constraints improves the FoM by a factor of 2500.\\ We should stress that even the results for the SDSS DR7 survey (with a relative FoM of 0.6) are promising, because they are a different way of probing the dynamics and therefore can be potentially useful in determining consistency of the underlying cosmological model. Clearly the addition of void ellipticities as an observable for parameter estimation increases our knowledge of the cosmological parameters in other cases. \subsection{Study of Possible Systematics} \label{ResultsIssues} While we have shown that our forecasted constraints are extremely promising, we have used order of magnitude calculations often based on first order results in semi-analytic models. By doing N-body simulations of large scale structure it is possible to replace these by more accurate calculations, and use it for estimating cosmological parameters. This would be the goal of future work in this direction. But is it possible that when such a rigorous analysis is carried out the constraints might get terribly degraded and not be interesting any more? The objective of this subsection is to address this concern by trying to list the major assumptions that would need to be replaced in a rigorous calculation, and trying to obtain a sense for how far these constraints might be degraded. We discuss the basic assumptions and explain how we might expect these factors to affect the forecasts.\\ \begin{figure*} \begin{center} \includegraphics[width=0.95\textwidth]{Fig4} \caption{~\textbf{Effects of Shot Noise: Sensitivity of one sigma constraints to the efficiency of the void finder:} Degradation of constraints due to low efficiency of the void finder with the constraints shown in Fig.~\ref{Constraints} assuming high efficiency from the near future (left panel) and futuristic data (right panel). HST and PLANCK constraints were used in all these plots.} \label{VariationWithDelta} \end{center} \end{figure*} {(a)\it{Effects of Shot Noise on the Number estimate of Voids:}} \label{EffectsOfShotNoise} Our constraints are obviously dependent on our estimate of the number of voids that would be detected in a particular survey. Thus, regions of space which are not true voids but get misidentified as voids would cause a spurious enhancement of signal. Recall that voids have been defined as regions of space where the \emph{total matter density} is low (or minimum) but are identified by the low density of \emph{galaxies} which are biased baryonic tracers of the density field. The lack of direct knowledge of the dark matter density field is often addressed in the context of the Poisson Sample Model, where density contrast of galaxies is described as a Poisson point process with a mean density proportional to the dark matter density. Thus, there is a chance of identifying a region which has low density of galaxies but not dark matter as a void. Consequently, due to shot noise, one can only confidently infer a region of low galaxy density to be a void if the region is large relative to the average separation $l_{sep}(z)\sim (n_{gal}^{bg})^{-1/3}$ of visible galaxies at that redshift. This means that small voids might not really be voids, and the problem is exacerbated by the fact that the number of voids increases exponentially with smaller sizes of voids. A sophisticated treatment of this problem would associate a probability to describe the confidence of detection (for example as in ~\citet{2008MNRAS.386.2101N}) and incorporate that in the Likelihood. We use a rough model to estimate the importance of this effect by only choosing a minimum radius $R_{min}^{shot}(z,survey)$ of voids related to the $l_{sep}(z)$ as discussed before. A larger value of $A$ results in a larger values of $l_{sep}(z)$ which leads to a higher threshold for the minimum size of voids observed in the survey. Since the minimum radius of voids is set by the maximum of this survey dependent $R_{min}^{shot}$ and the survey independent $R_{min}^{Vinc}$ (Void in Cloud), this changes the numbers of voids strongly where $R_{min}^{shot}$ is much smaller than $R_{min}^{Vinc}$. We therefore compare the constraints for a pessimistic value of $A =2$ to the constraints obtained in Fig.~\ref{Constraints} with $A=1$. In Fig.~\ref{VariationWithDelta}, we show the Fisher forecasts for values of $\Delta$ assuming the same value $\sigma_\epsilon=0.1$ for both cases. The red ellipse with open circles show the constraints from Voids (SDSS) + HST + PLANCK in the left panel, and Voids (EUCLID) + PLANCK + HST (right panel) for $A=1,$ while the open green squares show the same constraints if $A=2$. When additionally, supernovae data is used: on the left panel we have DES SNe + HST +PLANCK + SDSS Voids, while on the right panel we use LSST SNe + HST + PLANCK + EUCLID Voids. The solid, thin black ellipse shows these constraints for $A=1$, while the solid thick blue ellipse show these constraints for $A=2$. For reference, we use the black "+" to show the constraints from DES SNe+ PLANCK + HST on the left panel, and LSST SNe + PLANCK + HST on the right panel. Clearly, while the constraints change, there is no severe degradation due to shot noise for the case based on DR7 survey, while this is somewhat important for the case based on EUCLID. We summarize the degradation in terms of a relative FoM in Table.~\ref{table:FoM}. \begin{figure*}[!tp] \begin{center} \includegraphics[width=0.9\textwidth]{Fig5} \caption{{\textbf{Impact of Bias}}: Degraded constraints on voids due to marginalization over a linear scale independent bias compared to constraints shown in Fig.~\ref{Constraints} for data from the near future (left panel) and futuristic data (right panel). HST and PLANCK priors were considered for all of these plots. } \label{BiasFig} \end{center} \end{figure*} {(b) \it{Bias:}} Since the observations pertain to galaxies rather than the dark matter distribution, we have no direct knowledge of the dark matter distribution even though the galaxy distribution and dark matter distribution are correlated. The qualitative understanding of the situation is that galaxies form due to the collapse of baryons into gravitational potential wells of collapsed dissipation-less dark matter. The simplest popular idea of linear scale independent bias models this by assuming that locally, the dark matter density contrast $\delta_g$ is proportional to the the total matter density contrast $\delta_m$, and the constant of proportionality is called the bias $b$. Bias different from unity affects our forecasts in two ways: (i) first, the Lagrangian radius of the void is estimated incorrectly as a function of $\delta_g$ rather than $\delta_m$. This leads to the use of a variance $\sigma_R$ on the incorrect scale, and second (ii) since we use the probability distribution of the eigenvalues conditioned on the density contrast of the voids, this changes the distribution of the eigenvalues. To address the issue of bias, we recalculate the forecasts by adding an extra parameter, the bias $b$ to our set of cosmological parameters and marginalize over $b$ as a nuisance parameter. The Fisher constraints for the near future are presented in the left panel of Fig.~\ref{BiasFig}, while the right panel shows the constraints for the far future. In both cases, the red open circles show the constraints of Voids + PLANCK + HST from the upper panel of Fig.~\ref{Constraints}, while the solid thin black line shows the constraints from Voids +PLANCK + HST +SNE, where it was assumed that $b=1$. The green open squares show the corresponding constraints for Voids + HST + PLANCK, and the thick blue solid ellipses show the constraints for Voids + HST +PLANCK + SNE, when the bias is marginalized over. \\ \begin{figure*}[t] \begin{center} \includegraphics[width=0.95\textwidth]{Fig6} \caption{{\textbf{Sensitivity of Fisher Constraints with respect to the prescription of Void Selection:}} Comparison of the constraints (red open circles) from voids shown in Fig.~\ref{Constraints} with other prescriptions. PLANCK and HST priors were used in all these plots. The other prescriptions lead to better constraints} \label{Degradation_Algorithm} \end{center} \end{figure*} {(c) \it{Void Selection Prescription}} While the eigenvalues of the void ellipsoid are expected to trace the eigenvalues of the tidal ellipsoid, the eigenvalues themselves are stochastic quantities and the connection to theory comes from studying the distribution of these eigenvalues. Hence it is important to select a set of voids from the data that will accurately reflect the theoretical distribution computed. As discussed in ~\citet{2008MNRAS.387..933C}, the void finders available use different methods to identify voids, and these result in different definitions of voids. A number of these void finders are based on demarcating contiguous regions of space of different shapes through some variant of a clustering algorithm, while other void finders like ~\citet{2009arXiv0906.4101L} identify voids from a density field smoothed at a particular length scale. On the theoretical side, we can compute the probability distribution of the eigenvalues of the tidal tensor analytically through the Doroshkevich formula Eqn.~\ref{DKV}, which we use in the computations here, which is the distribution valid at all points in space rather than at voids in particular. One may also compute the distribution of the eigenvalues (i) for a void of size $R$ identified with the size of the fluctuation at shell crossing as shown in subsections ~\ref{distribution_eigenvalues} and ~\ref{evolution_eigenvalues}, or (ii) at the minima of the density field when smoothed at a particular length scale (eg. see Appendix B of ~\citet{2009arXiv0906.4101L}). Both of these are not analytic estimates, but they can used to construct samples of the eigenvalue distributions using Monte Carlo methods and lend themselves naturally to use with the two classes of void finders respectively. The use of computationally intensive Monte Carlo is beyond the scope of this paper based on Fisher estimates. Instead we use the analytic Doroshkevich formula which was shown to be close to both of these distributions, but this requires us to identify the set of voids that correspond to the voids obtained by smoothing the density field at a particular Lagrangian scale $R_{\text{Smooth}}$. If we find a set of voids at a particular redshift of a set of different sizes, how can we identify what smoothing scale these voids correspond to? Given a set of point particles in space, we understand the action of smoothing: it tends to homogenize the field at scales below the smoothing scale. Thus, one may expect that on smoothing by a scale $R_{\text{Smooth}}$, one will be left with voids with distribution such that there are few voids of size below $\approx R_{\text{Smooth}},$ while the smoothing operation may slightly modify the shapes and sizes voids of larger size. At a particular redshift, the probability of forming large voids is much smaller than forming smaller voids. Consequently, the distribution of sizes of voids when the density field is smoothed to a scale $R_{\text{Smooth}}$, should be peaked at $\sim R_{\text{Smooth}}.$ From simulations used in ~\citet{2009arXiv0906.4101L}, it appears that the distribution of the number of voids with radius $R$ in a density field smoothed by a filter of size $R_{ \text{Smooth}}$, is peaked at $R \approx 4 R_{\text{Smooth}}$ and falls off rapidly above that. While this inspired our choice for identification of voids, it is important to keep in mind that the distribution depends on the cosmological parameters through $\sigma_{lin}(R,z)$. Consequently using an inaccurate selection criterion for voids can introduce biases in parameter estimation, and the correct prescription may also change the errors and constraints. In order to get a sense for how severely the constraints might be degraded when this is done, we compute the constraints for three different prescriptions of identification the set of voids and compare how far the constraints are degraded in different cases that suggest themselves. From the right panel of Fig.~\ref{GrowthDerivs}, we see that the distribution gets broader for larger values of $\sigma_R$. Since this corresponds to lower theoretical predictability, we should expect the parameter constraints to get degraded as the filtering scale $R$ becomes smaller. On the other hand, this will lead to a larger number of voids since there are many more smaller voids.\\ One may expect that when the density field is smoothed at $R_{\text{Smooth}}$, a non-negligible fraction of the voids have radii between $R_{\text{Smooth}}$ and $4R_{\text{Smooth}}$. We can therefore use a different limit $R = R_{\text{Smooth}}$ in accordance with our calculations using the generalized excursion set formalism in subsection. ~\ref{distribution_eigenvalues}. Finally, if we assume that all voids larger than a particular smoothing scale would be found, we can take $R_{\text{Smooth}}= Min(\{R\})$ found in that redshift bin. This is similar to the method adopted by ~\citet{2007arXiv0704.0881L}. The corresponding constraints are shown in Fig.~\ref{Degradation_Algorithm}. The red open circles show the constraints shown in Fig.~\ref{Constraints} for the prescription where $R_{ \text{Smooth}}= R/4,$ while the blue asterisks show the constraints obtained for the case where $R_{\text{Smooth}} = R$, and the open green squares show the constraints for the case where $R_{\text{Smooth}}=Min(\{R\})$. \\ \begin{figure*} \begin{center} \includegraphics[width=0.95\textwidth]{Fig7} \caption{\textbf{Sensitivity of Fisher Constraints with $\sigma_\epsilon$:} Comparison of constraints from voids (open red circles) with $\sigma_{\epsilon}=0.1$ shown in Fig.~\ref{Constraints} with pessimistically degraded constraints from voids due to a larger $\sigma_{\epsilon}=0.4$ for data in the near future (left panel) and in the far future (right panel). Despite the degradation, the constraints are still interesting. HST and PLANCK priors were used in all the plots. } \label{VariationWithSigmaEpsilon} \end{center} \end{figure*} {(d)\it{Sensitivity to Error Levels}} As discussed before, in our method of forecasting for Fig.~\ref{Constraints}, we have used a Gaussian Likelihood with an error $\sigma_{\epsilon}=0.1$ assuming that its order was set by the uncertainty of measuring the void size which was limited by the size of the void shell (if $\Delta \approx 0.4$). Indeed, this seems larger than the values of the error levels computed in section 5.3.2 of ~\citet{2009arXiv0906.4101L}. Further, in our analysis, we have assumed that the ellipticities of the mass tensor of voids are perfect tracers of ellipticity of the tidal tensor. More realistically, there would be some scatter around the correlation as shown in section 5.2 of ~\citet{2009arXiv0906.4101L}. It is quite possible that scatter of this kind, or the assumptions that we have made might increase the level of error bars on $\epsilon$ quantitatively. Therefore, we investigate the sensitivity of the constraints to the value of $\sigma_{\epsilon},$ the error to which the ellipticity was assumed to be measured.\\ We show these constraints in Fig.~\ref{VariationWithSigmaEpsilon}, where the contours with red open circles show the constraints using Voids + PLANCK + HST shown in the upper panels of Fig.~\ref{Constraints} with $A=1$ and $\sigma_{\epsilon}=0.1$, while the open green squares are the constraints where $\sigma_{epsilon}$ has been increased to $0.4$. The solid lines show the constraints where the constraints are estimated with simultaneous use of the SNe data, ie. DES SNE for the left panel and LSST SNe for the right panel. The thin black solid line is for $\sigma_{\epsilon}=0.1$, while the thick blue solid line is for $\sigma_{\epsilon}=0.4$. The contours in black "+" symbols show the constraints from SNe + PLANCK + HST for reference. \begin{table*}[!h] \caption{Relative Figure of Merit (FoM) for using voids} \label{table:FoM} \begin{tabular}{|c|c|c|c|c|} \hline \multicolumn{1}{|c|}{}&\multicolumn{2}{|c|}{SDSS+DES+HST+PLANCK} &\multicolumn{2}{|c|}{EUCLID+LSST+HST+PLANCK}\\ \hline Parameters & Voids+CMB+HST & Voids + CMB+ HST +SNE& Voids+CMB+HST& Voids+CMB+HST+SNE\\ \hline $A=1,\sigma=0.1$ & 1.2 & 16.8 & 8.8& 331.0 \\ \hline $A=2,\sigma =0.1$ & 0.6 & 13.3 & 0.5 & 21.3 \\ \hline $A =1,\sigma=0.4$ & 0.5 & 7.7 & 0.7 & 27.6 \\ \hline Marginalized over $b$ & 0.2 & 3.3 & 0.2 & 104.5 \\ \hline $R_{\text{Smooth}}= Min(\{R\})$ & 6.1 & 24.9 & 3.6 & 73.0 \\ \hline $R_{\text{Smooth}}= R$ & 6.1 & 24.7 & 4.8 & 85.2 \\ \hline \end{tabular} \end{table*} \section{Summary and Discussions} The growth of cosmic structures with time depends on the background cosmology. Consequently, the growth of structures have been used to constrain the parameters of the background cosmology. Traditionally, the measures of growth used have characterized the growth of the volume of fluctuations. However, since the fluctuations are not individually isotropic, there is further information about the cosmology in the growth of asymmetry of the structures which could be extracted from its shape. Such a quantity parametrizing the shape of voids and its evolution was studied in ~\citet{2007PhRvL..98h1301P,2007arXiv0704.0881L}. The basic idea is that void shapes can be approximated as ellipsoidal structures, and relative sizes of the principal axes can be used as tracers of functions of eigenvalues of the tidal ellipsoid. In a spectroscopic survey, all three axes of the void ellipsoid may be measured, and thus asymmetry parameters which describe the shape of the ellipsoid are related to the quantities involving the eigenvalues of the tidal tensor, which depend on the background cosmology through the linearly extrapolated variance in fluctuations. Such spectroscopic surveys have been planned for studying large scale structure using traditional methods; thus the use of shapes does not necessarily require new surveys, but allows one to leverage data in an additional way. {\citet{2009arXiv0906.4101L} show that recovering the the tidal ellipticity of voids to high precision is indeed feasible. To do so, they identify voids and characterize the void tidal ellipticity using the simulated galaxy positions derived from a numerical simulation. These derived ellipticities are then compared to the tidal ellipticity of the complete displacement field given by the simulation. In this paper, we study the constraints on dark energy parameters from future surveys in terms of Fisher forecasts. The likelihood is a strong function of the linearly extrapolated variance of fluctuations at the redshift of the void at the scale of the Lagrangian size of the void. Since voids expand in comoving coordinates, their Lagrangian size is smaller than their observed (comoving) size, and this corresponds to a larger variance. variance at a smaller scale than the observed void size. We assume an error model with Gaussian noise on the measured ellipticity of the voids, and an arbitrarily assumed error on the ellipticity. We provide explicit formulae for Fisher matrices, and an estimate of the number of voids expected to be found from planned future surveys using semi-analytic methods. By comparing these Fisher constraints using void shapes from these surveys to the traditional constraints from other measures, we find this method to be promising: the constraints are quite competitive with traditional probes in the near future and combining the constraints with supernovae data improves the DETF Figure of Merit for the supernovae data by a factor of about ten. For futuristic data, we find that the constraints are close to ten times better than supernovae data, and combining with supernovae data, we can improve the FoM by a factor of a few hundred. We have used the Doroshkevich formula for the ellipticity throughout, but it has been shown~\citep{2009arXiv0906.4101L} that the distribution of ellipticity for a minima in the density field is slightly different. In actual parameter estimation, we will have to account for this. We shall also have to use the scatter in the correlation of the ellipticity of the void ellipsoid with the real shape of the tidal tensor as obtained from specific void identification algorithms. An issue we have not addressed here is the ellipticity of voids that can be generated due to redshift distortions ~\citep{1995ApJ...452...25R,1996ApJ...470..160R} which would have to be modeled to obtain unbiased parameter constraints from voids. The Fisher constraints are computed using simple models of dynamics and a likelihood. For estimation of parameters, each of these would need to be computed precisely. In the subsection ~\ref{ResultsIssues}, we discuss some of the main sources of errors and ambiguities in our forecasts. We indicate how more rigorous, though computationally intensive calculations may be devised. We attempt to estimate how the parameter constraints might be affected by these more rigorous methods. While the constraints are often weakened, they still remain at least competitive with other constraints in the near future and the far future. In the case of futuristic surveys, addition of the void ellipticity to other constraints result in an improvement of the FoM by a factor of at least a hundred, in spite of degradation due to additional systematics. We therefore feel that our study makes a strong case for pursuing this idea in greater detail. \section{Acknowledgments} We would like to thank G.~Lavaux for many useful discussions and sharing insights from his results that motivated some choices in this paper. RB would like to thank W.M.~Wood-Vasey for discussions about LSST supernovae. The authors would like to thank the California Institute of Technology for hospitality during which part of this work was done. The authors acknowledge financial support from NSF grantAST 07-08849.
\section{Introduction} \label{intro} Edge operations on graphs such as edge rotations, or switchings, have been widely used in various contexts: to provide a notion of distance between graphs, for example by measuring the number of moves required to get from one graph to another, see for example \cite{Ch-Sa-Zo, Go-Sw, Ja, Jo}; to provide algorithms to transform one planar graph or tree into another, see \cite{Ai-Re}, and \cite{Bo-Hu} for an overview; as a tool in random graph theory to estimate probabilities, as exploited in \cite{Mc-Wo, Wo}. We may view such operations as part of a \emph{dynamic} theory, whereby a graph evolves in discrete steps into a different configuration. In this article we are motivated by geometric considerations to study how a graph may evolve into one that is as regular as possible, or into a prescribed configuration. In smooth geometry, curvature can be detected by the convergence or divergence of nearby geodesics. The Gauss-Bonnet Theorem gives an expression for the total curvature of a compact surface in terms of its Euler characteristic. On a planar graph $\Gamma$ there is a notion of \emph{combinatorial curvature} introduced by Y. Higuchi \cite{Hi}, given by the function $\Phi$ defined at each vertex $x$ by the formula: $$ \Phi (x) = 1 - \frac{d(x)}{2} + \sum_{\sigma} \frac{1}{|\sigma |}\,, $$ where $d(x)$ denotes the degree of $x$, that is the number of edges incident with $x$, and where the sum is taken over all polygons $\sigma$ incident with $x$, with $|\sigma |$ representing the number of sides of $\sigma$. For a finite connected planar graph, the total curvature is given by $\sum_x\Phi (x) = 2$. As a variant, on an arbitrary graph, we can take the integer-valued function $K$ defined at each vertex $x$ by $K(x) = 2 - d(x)$. Up to a multiple of $2$, this may be viewed as an approximation of $\Phi$ on a sparse graph, that is a graph with few connections. Then we quickly deduce an analogue of the Gauss-Bonnet Theorem: $\sum_{x} K(x) = 2 \chi (\Gamma )$, where the sum is taken over all vertices and where $\chi (\Gamma )$ is the Euler characteristic given by $n - e$ with $n$ the number of vertices and $e$ the number of edges of $\Gamma$. The problems we address here in a combinatorial setting are akin to the problem of uniformization and that of prescribing the scalar curvature on a surface, as described for example, in \cite{Au}. An edge slide is an operation whereby we slide one end of an edge along another edge, so a triple of the form $x \sim y \sim z$ with $x \not\sim z$ becomes either $x \sim z \sim y$ or $ y \sim x \sim z$, where the notation $x\sim y$ means that $x$ and $y$ are adjacent vertices. An edge slide is a purely local operation which preserves the connectedness and the Euler characteristic of a graph. A basic question is to know whether one can transform one graph into another by edge slides. Given two connected simple graphs on the same number of vertices and edges, this was shown to be possible by M. Johnson \cite{Jo}. His proof is non-constructive in that it doesn't provide an algorithm to carry out the required sequence of edge slides. Our first result provides a constructive proof of a slightly more general theorem. Let $\Gamma = (V, E)$ and $\Sigma = (W, F)$ be two finite connected simple graphs with the same number of vertices and edges and let $\psi : V \rightarrow W$ be a bijection between the vertex sets; we will refer to $\Gamma$ as the \emph{initial configuration} and $\Sigma$ as the \emph{prescribed configuration}. Then we show there is a combination of edges slides on $\Gamma$ to produce a new graph $\wt{\Gamma} = (V, \wt{E})$ such that $\psi : \wt{\Gamma} \rightarrow \Sigma$ is an isomorphism of graphs. Furthermore, we provide an algorithm to carry out the sequence of edges slides. A first step is to show that any single move of an edge which preserves connectedness can be achieved by edge slides, furthermore we provide a specific algorithm to do this. Our strategy is then to select the vertex $y$ of smallest degree in $\Sigma$ and to increase (if necessary) the degree of the corresponding vertex $x$ in $\Gamma$ until it has degree $n-1$. This is done by taking a spanning tree and simultaneously evolving both the spanning tree and the graph by edge slides. We then have to remove edges incident with $x$ in an appropriate way until its degree coincides with that of $y$. Our objective is to remove each of $x$ and $y$ from $\Gamma$ and $\Sigma$, respectively and so to reduce the problem to graphs of successively smaller size. A difficulty that may arise is that the complements of $x$ and $y$ may not be connected. This is overcome by first making a judicious choice of moves in both graphs. Recall that a degree sequence is a list of non-negative integers $(d_1, \ldots , d_n)$. The theorem of Erd\H{o}s and Gallai \cite{Er-Ga} gives the conditions when such a sequence has a realization as the degrees of the vertices of a simple graph. Two non-isomorphic graphs may have the same degree sequence. A regular graph is one with all vertices of the same degree. The problem of generating regular graphs, or more generally, graphs of restricted degree sequences, is an important aspect of the theory of random graphs, see \cite{Mc-Wo, Wo}. If we view the degree sequence as a combinatorial analogue of the metric and take $K$ as above for the curvature, then edge sliding exchanges degrees from one vertex to a neighbouring vertex and as such, appears as an analogue of the evolution of a metric by its curvature. We introduce a natural energy functional ${\mathcal E}(\Gamma )$ associated to a graph $\Gamma$ which measures its discrepancy from being regular. We then describe an algorithm to regularize $\Gamma$ by edge slides, which at each step decreases ${\mathcal E}$. At the end of the algorithm, the graph is in what we call an \emph{almost regular configuration}, which is as close to being regular as possible given the number of vertices and edges. Finally, we allow the creation of vertices and edges in such a way as to preserve the connectivity and the Euler characteristic of a connected simple graph. We then provide an algorithm for such a graph $\Gamma$ to evolve into a prescribed configuration $\Sigma$ with the same Euler characteristic, where we no longer require $\Gamma$ and $\Sigma$ to have the same number of vertices and edges. \section{Notation and terminology} \label{sec:slides} Let $\Gamma = (V, E)$ be a simple graph, where $V$ denotes the vertices and $E$ the edges. Thus we do not allow double edges or loops. We will write an edge $\varepsilon \in E$ in the form $\varepsilon = \ov{xy}$ when we want to indicate the end points $x,y \in V$, otherwise we also write $x \sim y$ to indicate that $x$ and $y$ are joined by an edge and we will say that $x$ and $y$ are \emph{adjacent} or are \emph{neighbours}. We set $d(x)$ to be the number of edges incident with the vertex $x$. Then $\sum_{x \in V} d(x) = 2e$, where $e = |E|$ is the cardinality of $E$. We now define the operation of sliding. Let $x,y,z \in V$ be three distinct vertices such that $x \sim y \sim z$ and $x\not\sim z$. We call such a configuration of vertices a \emph{triple}. Then a \emph{(simple) slide along the edge $\ov{yz}$ with pivot $x$} creates the new configuration $x\sim z \sim y$. In most situations, we shall simply say that we slide the edge $\ov{xy}$ to $\ov{xz}$. Similarly, a slide along the edge $\ov{xy}$ with pivot $z$ creates the new configuration $y \sim x \sim z$. Both moves are possible since $x \not\sim z$. In general, we shall refer to the process of performing a sequence of edge slides as \emph{sliding}. \medskip \begin{center} \setlength{\unitlength}{0.254mm} \begin{picture}(400,72)(30,-106) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(40,-40)(40,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(40,-100)(100,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(200,-40)(200,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(360,-100)(420,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(120,-70)(160,-70)\special{sh 1}\path(160,-70)(154,-68)(154,-70)(154,-72)(160,-70) \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(40,-40){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(40,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(100,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(200,-40){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(200,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(260,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(360,-40){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(360,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(420,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\put(26,-46){\shortstack{$x$}} \special{color rgb 0 0 0}\put(26,-106){\shortstack{$y$}} \special{color rgb 0 0 0}\put(105,-106){\shortstack{$z$}} \special{color rgb 0 0 0}\put(186,-46){\shortstack{$x$}} \special{color rgb 0 0 0}\put(186,-106){\shortstack{$y$}} \special{color rgb 0 0 0}\put(265,-106){\shortstack{$z$}} \special{color rgb 0 0 0}\put(346,-46){\shortstack{$x$}} \special{color rgb 0 0 0}\put(346,-106){\shortstack{$y$}} \special{color rgb 0 0 0}\put(425,-106){\shortstack{$z$}} \special{color rgb 0 0 0}\put(295,-71){\shortstack{or}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(200,-40)(260,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(360,-40)(420,-100) \special{color rgb 0 0 0} \end{picture} \end{center} \medskip An edge slide always preserves connectedness of a graph and in the case of an edge slide along $\ov{yz}$ with pivot $x$, the degree at $y$ decreases by one and the degree at $z$ increases by one, while all other degrees remain the same. Note also that sliding is a reversible operation and on allowing the trivial slide which involves no move, we have an equivalence relation on the set of graphs with $n$ vertices and $e$ edges. More precisely, we shall say that two graphs $\Gamma$ and $\Sigma$ are \emph{slide-equivalent} if there exists a sequence of edge slides in $\Gamma$ which yields a new graph $\wt{\Gamma}$ such that $\wt{\Gamma}$ and $\Sigma$ are isomorphic, that is, there exists a bijection $\psi : V \rightarrow W$ from the vertices of $\wt{\Gamma}$ to those of $\Sigma$ such that $x \sim y$ if and only if $\psi (x) \sim \psi (y)$. We write a path joining two vertices $x$ and $u$ in the form $\sigma = [x:u]$, or if we want to indicate the vertices along the path, by $\sigma = [x:s_1\cdots s_k:u]$, where $x = s_1$, $u = s_k$ and $s_i \sim s_{i+1}$ for $i = 1, \ldots , k-1$, or finally by $\sigma = [s_1s_2 \cdots s_k]$ if we do not wish to specify the end points. Let $\sigma : [s_1s_2 \cdots s_k]$ be a path and let $y$ be a vertex not contained in $\sigma$. Suppose that $y \sim s_1$ and $y \not\sim s_i$ for $i = 2, \ldots , s_k$. Then we can perform consecutive slides of the edge $\ov{ys_1}$, first to $\ov{ys_2}$, then to $\ov{ys_3}$ and so on, until we reach $\ov{ys_k}$. In this case we shall say that we \emph{slide the edge $\ov{ys_1}$ along the path $\sigma$ to the edge $\ov{ys_k}$}. Suppose, given the path $\sigma$ as above, that $y \sim s_i$ and that $y \not\sim s_j$ for some $i,j \in \{ 1, \ldots , k\}$. Then whatever other connections exist between $y$ and the vertices of $\sigma$, it is clear that by a combination of slides we may achieve the configuration whereby $y\not\sim s_i$ and $y \sim s_j$, with all other connections unchanged. We call such a move a \emph{shuffle of $\ov{ys_i}$ to $\ov{ys_j}$ (along $\sigma$ with pivot $y$)}. Shuffling is a useful composite move that we will employ in the next section. \medskip \begin{center} \setlength{\unitlength}{0.200mm} \begin{picture}(574,137)(43,-206) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(45,-140)(75,-170) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(75,-170)(115,-190) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(115,-190)(165,-200) \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(45,-140){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(75,-170){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(115,-190){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(165,-200){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(215,-190)(165,-200) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(215,-190)(255,-170) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(255,-170)(285,-140) \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(215,-190){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(255,-170){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(285,-140){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(165,-85){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(360,-140)(390,-170) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(390,-170)(435,-190) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(435,-190)(490,-200) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(490,-200)(545,-190) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(545,-190)(585,-170) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(585,-170)(615,-140) \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(360,-140){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(390,-170){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(435,-190){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(490,-200){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(545,-190){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(585,-170){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(615,-140){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(490,-85){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(165,-85)(45,-140) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(165,-85)(75,-170) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(165,-85)(115,-190) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(165,-85)(165,-200) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(165,-85)(215,-190) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(165,-85)(285,-140) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(490,-85)(360,-140) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(490,-85)(435,-190) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(490,-85)(490,-200) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(490,-85)(545,-190) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(490,-85)(585,-170) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(490,-85)(615,-140) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(310,-140)(340,-140)\special{sh 1}\path(340,-140)(334,-138)(334,-140)(334,-142)(340,-140) \special{color rgb 0 0 0}\put(170,-81){\shortstack{$y$}} \special{color rgb 0 0 0}\put(495,-81){\shortstack{$y$}} \special{color rgb 0 0 0}\put(65,-186){\shortstack{$s_i$}} \special{color rgb 0 0 0}\put(260,-186){\shortstack{$s_j$}} \special{color rgb 0 0 0}\put(380,-186){\shortstack{$s_i$}} \special{color rgb 0 0 0}\put(590,-181){\shortstack{$s_j$}} \special{color rgb 0 0 0}\put(215,-206){\shortstack{$s_{j-1}$}} \special{color rgb 0 0 0}\put(545,-206){\shortstack{$s_{j-1}$}} \special{color rgb 0 0 0} \end{picture} \end{center} \medskip \begin{minipage}[center]{12cm} \emph{\footnotesize The figure illustrates a shuffle of the edge $\ov{ys_i}$ to $\ov{ys_j}$; first slide $\ov{ys_{j-1}}$ to $\ov{ys_j}$, then $\ov{ys_{j-2}}$ to $\ov{ys_{j-1}}$ and so on.} \end{minipage} \section{Discrete evolution to a prescribed configuration} In this section we show by producing an algorithm, that any two connected simple graphs with the same number of vertices and edges are slide-equivalent. In fact we establish a slightly stronger result, which allows us to choose arbitrarily the bijection between the vertices. This is expressed by the following theorem. \begin{theorem} \label{thm:prescribed} Given two connected simple graphs $\Gamma$ and $\Sigma$ each with the same number of vertices and edges. Let $\psi$ be a bijection between the vertices. Then there exists a combination of edge slides of $\Gamma$ to produce a graph $\wt{\Gamma}$ such that $\psi : \wt{\Gamma} \rightarrow \Sigma$ is an isomorphism of graphs. \end{theorem} As the first part of our strategy to prove this Theorem, we show that any move of an edge which preserves connectivity can be achieved by a combination of slides. Furthermore, we provide a specific algorithm to do this. First we have a preliminary lemma. \begin{lemma} \label{lem:paths} Let $\Gamma = (V, E)$ be a connected simple incomplete graph. Let $\varepsilon = \ov{uv} \in E$ be an edge and let $x,y \in V$ be unconnected vertices such that the graph obtained from $\Gamma$ by removing $\varepsilon$ and adding $\ov{xy}$ is connected. Then, up to labelling of $u$ and $v$, there exist paths from $x$ to $u$ and from $y$ to $v$ not containing the edge $\ov{uv}$. \end{lemma} \begin{proof} Since $\Gamma$ is connected, there is a path from $x$ to the edge $\varepsilon$. Stop the path at the first vertex of $\varepsilon$ encountered and call this vertex $u$. This gives a path from $x$ to $u$ not containing $\varepsilon$. Similarly, $y$ is connected to either $u$ or $v$ by a path not containing $\varepsilon$. If $y$ is connected to $v$, then we are done. Suppose on the other hand, $y$ is connected to $u$. Then by the same argument, since the graph $\wt{\Gamma}$ obtained by removing $\varepsilon$ and adding $\ov{xy}$ is connected, there is a path in $\wt{\Gamma}$ from $v$ to either $x$ or $y$ not containing $\ov{xy}$. If this is to $y$, we are done. If it is to $x$, then on relabeling $u$ and $v$, we are able to obtain paths from $x$ to $u$ and from $y$ to $v$ not containing $\ov{uv}$. \end{proof} \begin{lemma} \label{lem:moves} Any move of an edge which preserves connectedness is a combination of slides. \end{lemma} \begin{proof} Suppose we move $\varepsilon = \ov{uv}$ to $\ov{xy} \ ( x \not\sim y$ in $\Gamma$). Let $\sigma = [x:s_1\cdots s_k:u]$ be the shortest path from $x$ to $u$ not containing $\ov{uv}$ and let $\tau = [y:t_1 \cdots t_l:v]$ be the shortest path from $y$ to $v$ not containing $\ov{uv}$. Up to labeling, by Lemma \ref{lem:paths}, these exist. We have a number of cases to consider. \medskip \begin{center} \setlength{\unitlength}{0.254mm} \begin{picture}(400,82)(30,-116) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(40,-40)(65,-65) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(65,-65)(95,-80) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(95,-80)(175,-95) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(175,-95)(210,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(210,-100)(250,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(250,-100)(290,-95) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(290,-95)(365,-70) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(365,-70)(395,-55) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(395,-55)(420,-40) \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(40,-40){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(65,-65){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(95,-80){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(175,-95){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(210,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(255,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(365,-70){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(395,-55){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(420,-40){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(290,-95){\ellipse{4}{4}} \special{color rgb 0 0 0}\put(26,-46){\shortstack{$x$}} \special{color rgb 0 0 0}\put(60,-78){\shortstack{$s_2$}} \special{color rgb 0 0 0}\put(170,-111){\shortstack{$s_{k-1}$}} \special{color rgb 0 0 0}\put(210,-116){\shortstack{$u$}} \special{color rgb 0 0 0}\put(255,-116){\shortstack{$v$}} \special{color rgb 0 0 0}\put(290,-111){\shortstack{$t_{l-1}$}} \special{color rgb 0 0 0}\put(365,-86){\shortstack{$t_3$}} \special{color rgb 0 0 0}\put(95,-94){\shortstack{$s_3$}} \special{color rgb 0 0 0}\put(400,-66){\shortstack{$t_2$}} \special{color rgb 0 0 0}\put(425,-46){\shortstack{$y$}} \special{color rgb 0 0 0} \end{picture} \end{center} \medskip \noindent \emph{Case 1}: $x = u$; then $\sigma$ is the trivial path. (a) $u \not\in\tau$: then slide or shuffle $\ov{uv}$ to $\ov{xy}$ along $\tau$ (with pivot $u = x$). \medskip \begin{center} \setlength{\unitlength}{0.254mm} \begin{picture}(152,58)(40,-136) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(80,-80)(140,-80) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(110,-120)(80,-80) \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(140,-80){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(80,-80){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(110,-120){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(80,-80)(45,-120) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(140,-80)(170,-120) \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(170,-120){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(45,-120){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(45,-120)(110,-120) \special{color rgb 0 0 0}\put(105,-136){\shortstack{$u=x$}} \special{color rgb 0 0 0}\put(175,-131){\shortstack{$y$}} \special{color rgb 0 0 0}\put(40,-136){\shortstack{$v$}} \special{color rgb 0 0 0} \end{picture} \end{center} \medskip (b) $u \in \tau$: Suppose $u = t_i$; slide $\ov{vu}$ to $\ov{vy}$ along $[u:t_it_{i-1}\cdots t_1:y]$; this is possible since if $v \sim t_j$, $j < i$, then there would be a shorter path from $y$ to $v$. Now slide $\ov{vy}$ along $[v:t_lt_{l-1}\cdots t_i:x]$ to $\ov{xy}$. Once more, this is possible, since if there exists $j$, $j>i$, with $y \sim t_j$, then there would be a shorter path from $y$ to $v$. \medskip \begin{center} \setlength{\unitlength}{0.254mm} \begin{picture}(182,58)(35,-116) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(40,-100)(80,-60) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(80,-60)(120,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(120,-100)(160,-60) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(160,-60)(200,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(40,-100)(120,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(40,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(80,-60){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(120,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(160,-60){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(200,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\put(35,-116){\shortstack{$v$}} \special{color rgb 0 0 0}\put(115,-116){\shortstack{$u=x$}} \special{color rgb 0 0 0}\put(200,-116){\shortstack{$y$}} \special{color rgb 0 0 0} \end{picture} \end{center} \medskip \noindent \emph{Case 2}: $x = v$: (a) $u \in \tau$: Suppose $u = t_i$; then we just slide $\ov{vu}$ to $\ov{vy}$ along $[u:t_it_{i-1}\cdots t_1:y]$ (with pivot $x = v$). \medskip \begin{center} \setlength{\unitlength}{0.254mm} \begin{picture}(182,58)(35,-96) \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(40,-80){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(120,-80){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(200,-80){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(80,-40){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(160,-40){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(80,-40)(200,-80) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(80,-40)(40,-80) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(40,-80)(120,-80) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(120,-80)(135,-65) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(145,-55)(160,-40) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(40,-80)(100,-60) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(125,-50)(160,-40) \special{color rgb 0 0 0}\put(200,-96){\shortstack{$y$}} \special{color rgb 0 0 0}\put(110,-96){\shortstack{$v=x$}} \special{color rgb 0 0 0}\put(35,-96){\shortstack{$u$}} \special{color rgb 0 0 0} \end{picture} \end{center} \medskip (b) $u \not\in \tau$: If $y = s_i \in \sigma$, then clearly $x = v \not\in [y:s_is_{i+1}\cdots s_k:u]$, so we can shuffle edge $\ov{vu}$ to $\ov{vy} = \ov{xy}$. \medskip \begin{center} \setlength{\unitlength}{0.254mm} \begin{picture}(182,58)(35,-136) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(40,-120)(80,-80) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(40,-120)(120,-120) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(120,-120)(160,-80) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(160,-80)(200,-120) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(80,-80)(135,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(40,-120){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(120,-120){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(80,-80){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(160,-80){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(200,-120){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(150,-105)(200,-120) \special{color rgb 0 0 0}\put(35,-136){\shortstack{$u$}} \special{color rgb 0 0 0}\put(115,-136){\shortstack{$v=x$}} \special{color rgb 0 0 0}\put(200,-136){\shortstack{$y$}} \special{color rgb 0 0 0} \end{picture} \end{center} \medskip If $y\not\in \sigma$, then we shuffle edge $\ov{uv}$ to $\ov{uy}$ and then $\ov{uy}$ to $\ov{xy}$. \medskip \begin{center} \setlength{\unitlength}{0.254mm} \begin{picture}(182,58)(35,-116) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(40,-100)(80,-60) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(80,-60)(120,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(120,-100)(160,-60) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(160,-60)(200,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(40,-100)(120,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(200,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(160,-60){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(120,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(80,-60){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(40,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\put(35,-116){\shortstack{$u$}} \special{color rgb 0 0 0}\put(115,-116){\shortstack{$v=x$}} \special{color rgb 0 0 0}\put(200,-116){\shortstack{$y$}} \special{color rgb 0 0 0} \end{picture} \end{center} \medskip By symmetry, the cases $y = v$ and $y = u$ are dealt with similarly. \medskip \noindent \emph{Case 3}: $x \neq u,v$, $y \neq u,v$: If $x\sim v$, we first slide $\ov{xv}$ to $\ov{xy}$; then from Case 1, we can now slide $\ov{uv}$ to $\ov{xv}$. If $x\not\sim v$, then we first slide $\ov{uv}$ to $\ov{xv}$ and once more, from Case 1, we can slide $\ov{xv}$ to $\ov{xy}$. \end{proof} The algorithm to perform a move of an edge by slides is therefore as follows: \emph{first construct the shortest paths from $x$ to $u$ and from $y$ to $v$ not containing $\ov{uv}$ (interchanging the labels $u$ and $v$ if necessary); then, depending on the case that occurs, proceed as in the above proof.} By interchanging two vertices in a graph, we mean the following. Let $\Gamma$ be a graph and let $x,y$ be two vertices of $\Gamma$, then \emph{interchanging $x$ and $y$} gives the new graph $\wt{\Gamma}$ with all neighbours of $x$ now neighbours of $y$ and all neighbours of $y$ now neighbours of $x$; all other edges remain unchanged. The following lemma shows that interchanging vertices can be achieved by sliding. \begin{lemma} \label{lem:interchange} In any connected simple graph, interchanging two vertices $x$ and $y$ can be achieved by a combination of slides. \end{lemma} \begin{proof} \emph{Case 1}: $x \sim y$: If $z$ is a vertex such that $ z \sim x$ and $z \sim y$, then there is no slide needed. If $z \sim x$ and $z\not\sim y$, then we slide $\ov{zx}$ to $\ov{zy}$. \medskip \noindent \emph{Case 2}: $ x \not\sim y$: Let $\sigma = [x:s_1\cdots s_l:y]$ be the shortest path joining $x$ to $y$; clearly $l\geq 3$. If $z\not\in \sigma$ is a vertex such that $z \sim x$ and $z\not\sim y$, then we can shuffle $\ov{zx}$ to $\ov{zy}$. Similarly, if $z\not\in \sigma$ with $z\not\sim x$ and $z\sim y$, we can shuffle $\ov{zy}$ to $\ov{zx}$. There remain the edges $\ov{xs_2}$ and $\ov{s_{l-1}y}$ to deal with. We observe that since $\sigma$ has minimal length, for any $3\leq i \leq l$, we have $x \not\sim s_i\not\sim y$. If $l = 3$, there is nothing left to do. If $l \geq 4$, then by sliding, we have to move edge $\ov{xs_2}$ to $\ov{ys_2}$ and $\ov{s_{l-1}y}$ to $\ov{s_{l-1}x}$ to finish the problem. We do this as in the diagram below: \medskip \begin{center} \setlength{\unitlength}{0.230mm} \begin{picture}(542,77)(30,-136) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(35,-120)(55,-80) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(55,-80)(75,-70) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(115,-120)(105,-95) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(75,-70)(105,-95) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(140,-100)(165,-100)\special{sh 1}\path(165,-100)(159,-98)(159,-100)(159,-102)(165,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(195,-80)(215,-70) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(255,-120)(245,-95) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(215,-70)(245,-95) \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(35,-120){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(55,-80){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(75,-70){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(105,-95){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(115,-120){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(175,-120){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(195,-80){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(215,-70){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(245,-95){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(255,-120){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(175,-120)(255,-120) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(280,-100)(305,-100)\special{sh 1}\path(305,-100)(299,-98)(299,-100)(299,-102)(305,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(315,-120)(395,-120) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(385,-95)(315,-120) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(335,-80)(355,-70) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(355,-70)(385,-95) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(525,-95)(455,-120) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(510,-105)(535,-120) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(475,-80)(495,-95) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(475,-80)(495,-70) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(495,-70)(525,-95) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(420,-100)(445,-100)\special{sh 1}\path(445,-100)(439,-98)(439,-100)(439,-102)(445,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(315,-120){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(385,-95){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(335,-80){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(355,-70){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(395,-120){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(455,-120){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(475,-80){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(495,-70){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(525,-95){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(535,-120){\ellipse{4}{4}} \special{color rgb 0 0 0}\put(30,-136){\shortstack{$x$}} \special{color rgb 0 0 0}\put(115,-136){\shortstack{$y$}} \special{color rgb 0 0 0}\put(80,-71){\shortstack{$s_3$}} \special{color rgb 0 0 0}\put(110,-96){\shortstack{$s_{l-1}$}} \special{color rgb 0 0 0}\put(170,-136){\shortstack{$x$}} \special{color rgb 0 0 0}\put(255,-136){\shortstack{$y$}} \special{color rgb 0 0 0}\put(220,-71){\shortstack{$s_3$}} \special{color rgb 0 0 0}\put(250,-96){\shortstack{$s_{l-1}$}} \special{color rgb 0 0 0}\put(310,-136){\shortstack{$x$}} \special{color rgb 0 0 0}\put(395,-136){\shortstack{$y$}} \special{color rgb 0 0 0}\put(360,-71){\shortstack{$s_3$}} \special{color rgb 0 0 0}\put(390,-96){\shortstack{$s_{l-1}$}} \special{color rgb 0 0 0}\put(450,-136){\shortstack{$x$}} \special{color rgb 0 0 0}\put(535,-136){\shortstack{$y$}} \special{color rgb 0 0 0}\put(530,-96){\shortstack{$s_{l-1}$}} \special{color rgb 0 0 0}\put(40,-76){\shortstack{$s_2$}} \special{color rgb 0 0 0}\put(180,-76){\shortstack{$s_2$}} \special{color rgb 0 0 0}\put(320,-76){\shortstack{$s_2$}} \special{color rgb 0 0 0}\put(460,-76){\shortstack{$s_2$}} \special{color rgb 0 0 0}\put(500,-71){\shortstack{$s_3$}} \special{color rgb 0 0 0} \end{picture} \end{center} \medskip First, slide $\ov{xs_2}$ to $\ov{xy}$; now slide $\ov{s_{l-1}y}$ to $\ov{s_{l-1}x}$, and finally, slide $\ov{yx}$ to $\ov{ys_2}$. \end{proof} \begin{lemma} \label{lem:spanning} Let $T$ be a tree on $n$ vertices (and so $n-1$ edges). Fix any vertex, then there exists a combination of slides which give this vertex the degree $n-1$. \end{lemma} \begin{proof} Note that since sliding does not change the number of vertices or edges of a graph, it must preserve the property that a graph be a tree. Clearly the result is true for $n = 1,2$, so suppose $n \geq 3$. Fix a vertex $x$. If $d(x) = n-1$ we are done, so suppose $d(x)<n-1$. There must be at least two vertices of degree $1$, otherwise we would contradict the fact that the sum of degrees is equal to $2(n-1)$. Furthermore, there must be at least one such vertex which is not a neighbour of $x$. Let $y$ be a vertex of degree $1$ not equal to $x$ and not a neighbour of $x$. Let $z$ be its neighbour. Then we can shuffle $\ov{yz}$ to $\ov{yx}$ so increasing the degree at $x$ by one. Repeat this process until $d(x) = n-1$. \end{proof} \begin{lemma} \label{lem:increasing-degree} Let $\Gamma$ be a connected simple graph and let $x$ be any chosen vertex. Then by sliding we can arrange for $x$ to have degree $n-1$. \end{lemma} \begin{proof} Consider the graph $\Gamma_1 = \Gamma$. Fix a vertex $x$. If $d(x) = n-1$ we are done, otherwise, take a spanning tree $T_1$ of $\Gamma_1$. Now begin the operation of the proof of Lemma \ref{lem:spanning} by performing a simple slide of an edge of $T_1$ in order to increase the degree at $x$. If the graph permits (so as not to create a double edge), slide the same edge in $\Gamma_1$, otherwise do nothing. We now have a new tree $T_2$ spanning a new graph $\Gamma_2$ (maybe $\Gamma_1 = \Gamma_2$). Repeat this process, so that at each step we obtain a connected simple graph $\Gamma_i$ obtained from $\Gamma_{i-1}$ by an edge-slide, together with a spanning tree $T_i$, in such a way that $T_i$ is obtained from $T_{i-1}$ by performing the requisite simple slide to increase the degree of the designated vertex. We continue until the designated vertex has degree $n - 1$. \end{proof} \medskip \setlength{\unitlength}{0.254mm} \begin{picture}(484,124)(38,-162) \special{color rgb 0 0 0}\allinethickness{0.508mm}\path(40,-100)(160,-100) \special{color rgb 0 0 0}\allinethickness{0.508mm}\path(100,-40)(100,-160) \special{color rgb 0 0 0}\allinethickness{0.508mm}\path(100,-40)(130,-70) \special{color rgb 0 0 0}\allinethickness{0.508mm}\path(100,-40)(70,-70) \special{color rgb 0 0 0}\allinethickness{0.508mm}\path(100,-160)(130,-130) \special{color rgb 0 0 0}\allinethickness{0.508mm}\path(100,-160)(70,-130) \special{color rgb 0 0 0}\allinethickness{0.508mm}\special{sh 0.3}\put(160,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.508mm}\special{sh 0.3}\put(130,-70){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.508mm}\special{sh 0.3}\put(100,-40){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.508mm}\special{sh 0.3}\put(70,-70){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.508mm}\special{sh 0.3}\put(40,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.508mm}\special{sh 0.3}\put(70,-130){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.508mm}\special{sh 0.3}\put(100,-160){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.508mm}\special{sh 0.3}\put(130,-130){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.508mm}\special{sh 0.3}\put(100,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(160,-100)(130,-130) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(70,-130)(40,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(40,-100)(70,-70) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(70,-70)(95,-70) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(105,-70)(130,-70) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(70,-70)(70,-95) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(70,-105)(70,-130) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(70,-130)(95,-130) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(105,-130)(130,-130) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(130,-130)(130,-105) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(130,-95)(130,-70) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(130,-70)(100,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(180,-100)(200,-100)\special{sh 1}\path(200,-100)(194,-98)(194,-100)(194,-102)(200,-100) \special{color rgb 0 0 0}\allinethickness{0.508mm}\path(220,-100)(340,-100) \special{color rgb 0 0 0}\allinethickness{0.508mm}\path(280,-40)(280,-160) \special{color rgb 0 0 0}\allinethickness{0.508mm}\path(280,-40)(250,-70) \special{color rgb 0 0 0}\allinethickness{0.508mm}\path(280,-160)(310,-130) \special{color rgb 0 0 0}\allinethickness{0.508mm}\path(280,-160)(250,-130) \special{color rgb 0 0 0}\allinethickness{0.508mm}\path(280,-100)(310,-70) \special{color rgb 0 0 0}\allinethickness{0.508mm}\special{sh 0.3}\put(340,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.508mm}\special{sh 0.3}\put(310,-70){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.508mm}\special{sh 0.3}\put(280,-40){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.508mm}\special{sh 0.3}\put(280,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.508mm}\special{sh 0.3}\put(250,-70){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.508mm}\special{sh 0.3}\put(220,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.508mm}\special{sh 0.3}\put(250,-130){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.508mm}\special{sh 0.3}\put(280,-160){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.508mm}\special{sh 0.3}\put(310,-130){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(250,-70)(220,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(280,-40)(310,-70) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(340,-100)(310,-130) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(250,-130)(220,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(250,-130)(250,-105) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(250,-95)(250,-70) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(250,-70)(275,-70) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(285,-70)(310,-70) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(310,-70)(310,-95) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(310,-105)(310,-130) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(310,-130)(285,-130) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(275,-130)(250,-130) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(360,-100)(380,-100)\special{sh 1}\path(380,-100)(374,-98)(374,-100)(374,-102)(380,-100) \special{color rgb 0 0 0}\allinethickness{0.508mm}\path(400,-100)(520,-100) \special{color rgb 0 0 0}\allinethickness{0.508mm}\path(460,-40)(460,-160) \special{color rgb 0 0 0}\allinethickness{0.508mm}\path(460,-100)(490,-70) \special{color rgb 0 0 0}\allinethickness{0.508mm}\path(460,-100)(430,-70) \special{color rgb 0 0 0}\allinethickness{0.508mm}\path(460,-160)(490,-130) \special{color rgb 0 0 0}\allinethickness{0.508mm}\path(460,-160)(430,-130) \special{color rgb 0 0 0}\allinethickness{0.508mm}\special{sh 0.3}\put(460,-160){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.508mm}\special{sh 0.3}\put(490,-130){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.508mm}\special{sh 0.3}\put(430,-130){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.508mm}\special{sh 0.3}\put(460,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.508mm}\special{sh 0.3}\put(400,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.508mm}\special{sh 0.3}\put(520,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.508mm}\special{sh 0.3}\put(490,-70){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.508mm}\special{sh 0.3}\put(430,-70){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.508mm}\special{sh 0.3}\put(460,-40){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(460,-40)(490,-70) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(430,-70)(400,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(400,-100)(430,-130) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(490,-130)(520,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(430,-70)(455,-70) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(465,-70)(490,-70) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(430,-70)(430,-95) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(430,-105)(430,-130) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(430,-130)(455,-130) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(465,-130)(490,-130) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(490,-130)(490,-105) \special{color rgb 0 0 0}\allinethickness{0.254mm}\dottedline{5}(490,-95)(490,-70) \special{color rgb 0 0 0} \end{picture} \medskip \begin{minipage}[center]{12cm} \emph{\footnotesize The figure illustrates the simultaneous evolution of a graph (bold and dotted lines) and its spanning tree (bold lines). At the first step, the top right hand edge of the spanning tree slides to increase the degree of the central vertex by one, while the graph remains unchanged. At the second step, both the graph and the tree evolve to increase the degree of the central vertex by one.} \end{minipage} \bigskip \noindent \emph{Proof of Theorem} \ref{thm:prescribed}. We will apply induction on the number $n$ of vertices in $\Gamma$. \medskip \noindent \emph{Step 1}: In $\Sigma$, consider a vertex $y$ of minimum degree $d_1$. Let $x = \psi^{-1}(y)$ be the corresponding vertex in $\Gamma$. By Lemma \ref{lem:increasing-degree}, by sliding, we can obtain a new graph $\wt{\Gamma}$ in which $x$ has degree $n - 1$. Let $C_1, \ldots , C_s$ be the connected components of the graph formed from $\wt{\Gamma}$ be deleting $x$ and all its incident edges. By connectivity of $\wt{\Gamma}$, each one of these components is connected to $x$ by one or more edges. Let $k_i$ be the number of vertices of $C_i$. Our aim is to decrease the degree of $x$ until it reaches $d_1$. To do this, we perform the following algorithm. \medskip \begin{center} \setlength{\unitlength}{0.254mm} \begin{picture}(216,136)(80,-145) \special{color rgb 0 0 0}\allinethickness{0.254mm}\put(102,-92){\ellipse{45}{35}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\put(160,-132){\ellipse{40}{25}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\put(245,-92){\ellipse{50}{35}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(165,-20)(90,-90) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(165,-20)(110,-95) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(165,-20)(160,-135) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(165,-20)(225,-95) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(165,-20)(245,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(165,-20)(255,-80) \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(90,-90){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(110,-95){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(165,-20){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(160,-135){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(225,-95){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(245,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(255,-80){\ellipse{4}{4}} \special{color rgb 0 0 0}\put(155,-21){\shortstack{x}} \special{color rgb 0 0 0}\put(80,-121){\shortstack{$C_1$}} \special{color rgb 0 0 0}\put(190,-141){\shortstack{$C_2$}} \special{color rgb 0 0 0}\put(265,-121){\shortstack{$C_3$}} \special{color rgb 0 0 0} \end{picture} \end{center} \medskip If $s \geq 2$, take an edge from $x$ to $C_1$ and use it to join $C_1$ to $C_2$. This maintains connectivity and by Lemma \ref{lem:moves}, this can be achieved by a number of slides. Now $C_1$ and $C_2$ have become one connected component. We repeat this process until \emph{either} we achieve the degree $d_1$, \emph{or} $d(x)>d_1$ and there exists just one connected component $C$ on removing $x$ and its incident edges. We claim that in the latter case there exists at least two non-adjacent vertices in $C$. If $C$ is complete, it has $n - 1$ vertices and $(n-1)(n-2)/2$ edges. Now the number of edges in $\wt{\Gamma}$ is $\geq d_1 + 1 + \frac{(n-1)(n-2)}{2}$. But, by removing $y$ from $\Sigma$ together with its incident edges, we see that $\Sigma$ has at most $d_1 + \frac{(n-1)(n-2)}{2}$ edges. But this is a contradiction, since all of our operations are achieved by sliding, which preserves the number of edges. Thus there exist two non-adjacent vertices in $C$ to which we can move an edge from $x$ to $C$. We now repeat this process until the degree $d_1$ is achieved at the vertex $x$. We now apply Lemma \ref{lem:interchange}. By interchanging vertices, we can arrange that the set of vertices that are neighbours of $x$ by the algorithm of Step 1, are exactly those vertices that are required to be neighbours in order that $x\sim y \Rightarrow \psi (x) \sim \psi (y)$. Write the new graph that results from Step 1 as $\wh{\Gamma}$. \medskip \noindent \emph{Step 2}: Let $\Gamma^{\prime}$ be the graph obtained from $\wh{\Gamma}$ by removing $x$ and its incident edges. We aim to apply induction, but $\Gamma^{\prime}$ may not be connected. Let $L_1, \ldots , L_t$ be its connected components. We will now connect them by slides. Let $L_i$ have $\ell_i$ vertices. If $L_1$ contains $\geq \ell_1$ edges, it contains a cycle and we can take one of its edges and join it in $\wh{\Gamma}$ to another connected component, $L_2$ say, without disconnecting $L_1$. We repeat this process. If finally, we have components $L_1, \ldots , L_r$ ($r\geq 2$) in $\Gamma^{\prime}$ each with $\ell_1, \ldots , \ell_r$ vertices and $\ell_1-1, \ldots , \ell_r - 1$ edges (where, for simplicity of notation, we maintain the symbols $\Gamma^{\prime}$ and $L_i$, even though these may have changed under the above operations), then the total number of edges is given by $(\ell_1-1)+ \cdots + (\ell_r - 1) + d_1 = n - 1 - r + d_1$. On the other hand, $|E| \geq d_1n/2$, since $d_1$ is the smallest degree in $\Sigma$ and $\Sigma$ and $\wh{\Gamma}$ have the same number of edges. Thus $$ d_1 \leq \frac{2(n-1-r)}{n-2}<2\,, $$ so that $d_1 = 1$. But this is a contradiction, since $r \leq d_1$. Thus $r = 1$ and $\Gamma^{\prime} = L_1$ is connected. Similarly, in the graph $\Sigma$, we perform the above operations in order that, on removing $y$ and its incident edges, the resulting graph $\Sigma^{\prime}$ is connected. Note that the operations of Step 2 have no effect on the neighbours of $x$ and $y$; these are preserved. As remarked in Section \ref{sec:slides}, sliding is a reversible operation, hence the slides performed in $\Sigma$ can be reversed. By induction, the theorem is true for the graphs $\Gamma^{\prime}$ and $\Sigma^{\prime}$ on $n-1$ vertices and the result follows. \hfill $\Box$ \section{Regularisation} \label{sec:reg} Given two positive integers $n$ and $e$, then by Euclidean division, there are unique $k$ and $r$ such that $2e = nk + r$ where $0 \leq r \leq n-1$. Then we have the following configuration of degrees: \begin{equation} \label{min-config} \begin{array}{rl} r & \left\{ \begin{array}{l} k+1 \\ \vdots \\ k+1 \end{array} \right. \\ n - r & \left\{ \begin{array}{l} k \\ \vdots \\ k \end{array} \right. \end{array} \end{equation} whose sum is $2e$. We can characterize such a configuration in the following way. For a degree sequence $(d_1, \ldots , d_n)$, define its \emph{energy} to be the quantity ${\mathcal E} = \sum_i d_i{}^2$. If each $d_i>0$, we shall say that the sequence is \emph{positive}. \begin{lemma} Given $e>0$, a positive degree sequence $(d_1, \ldots , d_n)$ with sum $2e$ minimizes ${\mathcal E}$ amongst all other positive sequences with sum $2e$ if and only if it has the form {\rm (\ref{min-config})} for some $k\geq 1$ and $r$ with $0 \leq r \leq n-1$. \end{lemma} \begin{proof} Among all sums $d_1{}^2 + \cdots + d_n{}^2$ with $d_1 + \cdots + d_n = 2e$ $(d_i >0)$, there is one which is minimal; let this be $S$. Let the sequence $(d_1, \ldots , d_n)$ realize $S$. We want to show this has the form (\ref{min-config}). If there are two integers $d_i$ and $d_j$ such that $d_i - d_j \geq 2$, then the sequence $(d_1, \ldots d_{j - 1}, d_j +1, d_{j+1}, \ldots d_{i-1}, d_i-1, d_{i+1}, \ldots d_n)$ has all terms $>0$, sum $2e$, but with sum of squares equal to $S + 2(d_j - d_i + 1) < S-1$, contradicting our hypothesis that $S$ is the minimum of ${\mathcal E}$ amongst positive sequences. This means that any two terms of the sequence $(d_1, \ldots , d_n)$ can either be equal or can differ by $1$, so they all have the form $t$ or $t+1$ for some positive integer $t$. But then the sum of the terms lies between $nt$ and $n(t+1)$ and $t$ must equal $k$. \end{proof} Call a graph $\Gamma = (V, E)$ such that $\max_{x,y \in V} |d(x)- d(y)| \leq 1$ an \emph{almost regular graph}. We now devise an algorithm for obtaining an almost regular graph from a given connected simple graph by sliding. \medskip \begin{center} \setlength{\unitlength}{0.254mm} \begin{picture}(404,84)(38,-102) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(40,-60)(120,-60) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(80,-20)(80,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(80,-20)(120,-60) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(120,-60)(80,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(80,-100)(40,-60) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(40,-60)(80,-20) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(100,-40)(60,-80) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(100,-80)(60,-40) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(200,-60)(280,-60) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(240,-20)(240,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(240,-20)(280,-60) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(280,-60)(240,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(240,-100)(200,-60) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(200,-60)(240,-20) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(220,-40)(260,-80) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(260,-40)(245,-40) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(235,-40)(220,-40) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(260,-80)(245,-80) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(235,-80)(220,-80) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(400,-20)(400,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(400,-20)(440,-60) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(400,-20)(360,-60) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(360,-60)(400,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(400,-100)(440,-60) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(380,-40)(400,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(420,-80)(400,-20) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(360,-60)(380,-60) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(390,-60)(395,-60) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(405,-60)(410,-60) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(420,-60)(440,-60) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(380,-40)(395,-40) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(410,-40)(420,-40) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(420,-80)(405,-80) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(390,-80)(380,-80) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(150,-60)(170,-60)\special{sh 1}\path(170,-60)(164,-58)(164,-60)(164,-62)(170,-60) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(310,-60)(330,-60)\special{sh 1}\path(330,-60)(324,-58)(324,-60)(324,-62)(330,-60) \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(120,-60){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(100,-40){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(80,-20){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(60,-40){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(40,-60){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(60,-80){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(80,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(100,-80){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(80,-60){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(240,-20){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(220,-40){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(200,-60){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(220,-80){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(240,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(260,-80){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(280,-60){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(260,-40){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(240,-60){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(400,-20){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(380,-40){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(360,-60){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(380,-80){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(400,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(420,-80){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(440,-60){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(420,-40){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(400,-60){\ellipse{4}{4}} \special{color rgb 0 0 0} \end{picture} \end{center} \medskip \begin{minipage}[center]{12cm} \emph{\footnotesize The figure shows the regularization according to the algorithm below of a graph on nine vertices and sixteen edges. Sliding occurs in such a way that degrees are exchanged from high incidence vertices to low incidence ones until an almost regular configuration is achieved. At each step two slides have occured.} \end{minipage} \bigskip \begin{theorem} Let $\Gamma = (V, E)$ be a connected simple graph. Then $\Gamma$ is slide-equivalent to an almost regular graph. \end{theorem} \begin{proof} If no two vertices have degrees which differ by at least $2$, then the graph is already in an almost regular configuration, so we suppose otherwise. We make the following algorithm: At each step, we take two vertices $x,y \in V$ such that $d(x) - d(y) \geq 2$. Let $\sigma = [x:s_1s_2\cdots s_k:y]$ be the shortest path from $x$ to $y$. If $k = 2$, then $x \sim y$ and because $d(x) - d(y) \geq 2$, there exists a vertex $a$ such that $a\sim x$, $a \not\sim y$ and $a \neq y$. Then we slide the edge $\ov{ax}$ along $\ov{xy}$, so decreasing the degree at $x$ by $1$, increasing the degree at $y$ by $1$ and leaving all other degrees the same. If $k \geq 3$, then $x\not\sim y$ and because $d(x) - d(y) \geq 2$, there are two distinct vertices $a,b$ such that $x\sim a$, $x \sim b$, $a \neq y\neq b$ and $a\not\sim y\not\sim b$. Since $\sigma$ is the shortest path from $x$ to $y$, we cannot have both $a$ and $b$ in this path; so suppose that $a \not\in \sigma$. Then we shuffle the edge $\ov{ax}$ along $\sigma$ until we connect $a$ to $y$. Now we see that after each step, the energy ${\mathcal E}$ strictly decreases and so this algorithm ends. In fact, for a connected simple graph, the degrees satisfy $1\leq d(x) \leq n-1$ at each vertex $x$, so there is only a finite number of possible degree sequences and consequently, only a finite number of possible values of ${\mathcal E}$. The graph we obtain when we finish has vertices all of whose degrees differ by at most $1$ and so is almost regular, as required. \end{proof} \begin{remark} The above theorem shows that any degree sequence of the form (\ref{min-config}) where $n-1\leq e\leq n(n-1)/2$ can be realized as the degree sequence of a connected simple graph. We simply construct any connected simple graph with $e$ edges and then apply the theorem to slide it into an almost regular configuration. \end{remark} \section{Expanding and collapsing a graph within its Euler class} \label{sec:exp} Theorem \ref{thm:prescribed} shows that any two connected simple graphs with the same number of edges and vertices are equivalent by edge slides. However, it would be useful to be able to change the number of vertices and edges and still establish equivalence under appropriate conditions. This may be important in random graph theory, where one is required to let $n \rightarrow\infty$ in an appropriate class of graphs, for example regular graphs, see \cite{Bo, Mc-Wo, Wo}. In this section, we show how it is possible to increase the number of vertices and edges without limit whilst preserving the Euler characteristic. Let $\Gamma = (V, E)$ be a finite graph. Recall that the Euler characteristic is the number $\chi (\Gamma ) = n - e$, where $n$ is the number of vertices and $e$ the number of edges. Note that if $\Gamma$ is connected and simple (no loops or double edges), then there are two extremes; a complete graph or a tree. For a complete graph $e = \frac{1}{2}n(n-1)$ and for a tree $e = n-1$, so that $$ n-1 \leq e \leq \frac{1}{2}n(n-1) $$ and we deduce that $$ \frac{n(3-n)}{2} \leq \chi \leq 1\,. $$ We now introduce a process for expanding and collapsing a graph. This involves two possible moves which are equivalent up to sliding. Furthermore, the moves preserve the Euler characteristic and the connectivity of the graph. \medskip \noindent \emph{Expanding}: (i) We add a new vertex $y$ to the graph which is attached to any given vertex $x$ by a new edge $\overline{xy}$ (so that $d(y) = 1$). (ii) To any edge $\overline{xz}$, we add a new vertex $y$ to its interior, so dividing the edge into two edges $\overline{zy}$ and $\overline{yx}$. \medskip \begin{center} \setlength{\unitlength}{0.254mm} \begin{picture}(497,86)(60,-140) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(80,-100)(160,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(60,-80)(80,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(80,-100)(60,-120) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(160,-100)(80,-140) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(160,-100)(180,-140) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(265,-100)(345,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(265,-100)(245,-80) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(265,-100)(245,-120) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(345,-100)(265,-140) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(345,-100)(365,-140) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(450,-100)(530,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(450,-100)(430,-80) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(450,-100)(430,-120) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(530,-100)(450,-140) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(530,-100)(555,-140) \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(80,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(160,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(265,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(345,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(365,-60){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(450,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(490,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\special{sh 0.3}\put(530,-100){\ellipse{4}{4}} \special{color rgb 0 0 0}\put(85,-96){\shortstack{$z$}} \special{color rgb 0 0 0}\put(170,-101){\shortstack{$x$}} \special{color rgb 0 0 0}\put(270,-96){\shortstack{$z$}} \special{color rgb 0 0 0}\put(355,-101){\shortstack{$x$}} \special{color rgb 0 0 0}\put(370,-66){\shortstack{$y$}} \special{color rgb 0 0 0}\put(455,-96){\shortstack{$z$}} \special{color rgb 0 0 0}\put(495,-96){\shortstack{$y$}} \special{color rgb 0 0 0}\put(540,-101){\shortstack{$x$}} \special{color rgb 0 0 0}\put(395,-106){\shortstack{or}} \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(200,-100)(230,-100)\special{sh 1}\path(230,-100)(224,-98)(224,-100)(224,-102)(230,-100) \special{color rgb 0 0 0}\allinethickness{0.254mm}\path(365,-60)(345,-100) \special{color rgb 0 0 0} \end{picture} \end{center} \medskip Both these operations preserve the Euler characteristic, since we add one edge and one vertex. Furthermore, if $z \sim x$, then after (i), we can slide $\overline{zx}$ to $\overline{zy}$, so that effectively, we have just added a vertex to the interior of the edge $\overline{zx}$. Conversely, if we add a vertex $y$ to the interior of $\overline{zx}$, then we can slide $\overline{zy}$ to $\overline{zx}$, leaving the isolated vertex $y$ such that $y \sim x$. Thus (i) and (ii) are equivalent by edge slides. \medskip \noindent \emph{Collapsing}: (iii) If $y$ is a vertex such that $d(y) = 1$, then delete $y$ and the edge that joins it to the rest of the graph. (iv) Let $y$ be a vertex such that $d(y) = 2$ and such that if $x\sim y$ and $z \sim y$ ($z \neq x$), then $x\not\sim z$; then delete the vertex $y$ together with the edges $\overline{zy}$ and $\overline{yx}$ and add the edge $\overline{zx}$. \medskip Both operations (iii) and (iv) remove one vertex and one edge and so leave the Euler characteristic the same. \begin{corollary} {\rm (of Theorem \ref{thm:prescribed})} Let $\Gamma$ and $\Sigma$ be two connected simple graph with $m$ and $n$ vertices, respectively, having the same Euler characteristic. Suppose $m \leq n$. Then we can expand $\Gamma$ using operations {\rm (i)} or {\rm (ii)} and perform a combination of edge slides to obtain a graph isomorphic to $\Sigma$. Equally, we can collapse $\Sigma$ using operations {\rm (iii)} or {\rm (iv)} and perform a combination of slides to obtain a graph isomorphic to $\Gamma$. \end{corollary} \begin{proof} The first part of the corollary is a direct consequence of Theorem \ref{thm:prescribed}: we add $(n-m)$ vertices and edges according to the rules (i) and (ii) (it doesn't matter where we place them); then we can slide the graph into a configuration that is isomorphic to $\Sigma$. For the second part, since the Euler characteristics of $\Gamma$ and $\Sigma$ coincide, we must have $$ \frac{m(3-m)}{2} \leq \chi \leq 1\,. $$ By induction, it suffices to show that the result holds for $m = n-1$. Substituting this into the above inequality, we have \begin{equation} \label{edges} \frac{(n-1)(4 - n)}{2} \leq \chi \leq 1 \Rightarrow f \leq \frac{(n-1)(n-2)}{2} + 1\,, \end{equation} where $f$ is the number of edges of $\Sigma$. Take any vertex $x$ of $\Sigma$ and let $C$ be the graph obtained from $\Sigma$ by removing $x$ and its incident edges. Let $C_1, \ldots , C_r$ be the connected components of $C$. Note that $r\leq d(x)$. If $d(x) >1$, then we take an edge joining $x$ to one of its neighbours and move it to two unconnected vertices in $C$. In the first instance we move such an edge to connect two disconnected components of $C$, if such exist. Note that there must always be two unconnected vertices in $C$, since if $C$ is complete, then it must contain $(n-1)(n-2)/2$ edges, contradicting the inequality (\ref{edges}). We repeat this process until $d(x) = 1$. Let $\widetilde{C}$ be the graph obtained by removing $x$ and the one edge incident with it. Then $\widetilde{C}$ must be connected (since $r \leq d(x)$). We now collapse the graph $\Sigma$ by removing the vertex $x$ of degree $1$ and the one edge incident with it, to obtain the graph $\widetilde{C}$ on $n-1$ vertices, with the same Euler characteristic $\chi$. By Theorem \ref{thm:prescribed}, by sliding in $\widetilde{C}$, we can obtain a graph on $n-1$ vertices isomorphic to $\Gamma$. \end{proof}
\section{Introduction} Geometric search problems have a long and rich history \cite{ae99,g00chapter,ms05}. In these problems, we are typically given a collection $S$ of $n$ geometric objects in $\R^d$ and asked to store them in a data structure so that we can efficiently answer queries about these objects. Until quite recently, the performance of these data structures was typically measured, at least theoretically, in terms of the worst-case (over all possible choices of $S$ and $q$) query time as a function of $n$. A worst-case query time of $O(\log n)$ is typically viewed as the gold standard for such search problems, at least in models of computation that only allow binary decisions. Somewhat more recently, researchers have begun studying geometric search problems under the lens of \emph{distribution-sensitivity}. In this setting, one assumes that there is a probability measure $D$ over the set of possible queries, and one attempts to optimize the expected query time when queries are distributed according to $D$. For example, given a planar triangulation $G$ and a distribution $D$ over $\R^2$, one can construct an $O(n)$ sized data structure that can determine the face of $G$ that contains any query point $q\in\R^2$. The expected query time of this structure is $O(H+1)$, where \begin{equation} H = \sum_{i} p_i\log (1/p_i) \enspace \eqlabel{entropy-faces} \end{equation} and $p_i$ denotes the probability that $q$ is contained in the $i$th face of $G$ \cite{acmr00,amm00,amm01a,amm01b,ammw07,i01,i04}. Information theory tells us that this result is optimal in any \emph{dichotomous} model of computation where any execution branching has at most 2 possible outcomes. More recently, Collette \etal\ \cite{cdilm08,cdilm09} have shown that a similar result holds for point location in any simple connected planar subdivision $G$. Note that this involves more than simply triangulating $G$ and applying the results for triangulations. Triangulating $G$ increases the number of faces and therefore also increases the value of $H$ in \eqref{entropy-faces}. On the other hand, we can not expect to always achieve a query time of $O(H+1)$ since this would imply, for example, an $O(1)$ query-time data structure for testing if a point is contained in a simple polygon. The result of Collette \etal\ is a data structure whose expected query time is $O(H^*+1)$ where $H^*$ is a lower-bound on the expected cost of any linear decision tree for point location in $G$ with queries distributed according to $D$. Dujmovi\'c \etal\ \cite{dhm09} study distribution-sensitive 2-sided 2-dimensional orthogonal range counting. They describe a data structure, called a biased range tree, that preprocesses an $n$ point set $S\subset\R^2$ and a query distribution $D$ over $\R^2$. A biased range tree uses $O(n\log n)$ preprocessing time and space and can answer 2-sided 2-dimensional orthogonal range counting queries over $S$ in expected time $O(H^*+1)$, where $H^*$ is a lower-bound on the expected cost of any comparison tree for 2-sided range queries over $S$ with queries distributed according to $D$. The results of Collette \etal\ \cite{cdilm08,cdilm09} and Dujmovi\'c \etal\ \cite{dhm09} use similar techniques to prove their optimality, but each requires their own \textit{ad hoc} arguments. For example, the point location results are achieved by finding near-minimum-entropy Steiner triangulations and then applying existing distribution-sensitive results for point location in triangulations. Biased range trees on the other hand, mix $k$-d trees, fractional cascading, and biased binary search trees to achieve their running time. In the current paper, we describe a general and low-overhead method of taking any $O(\log n)$ query-time data structure and making it distribution-sensitive. Let $\mathcal{P}:\R^d\rightarrow A$ be a query problem over $\R^d$ for which there exists a data structure $\mathcal{S}$ that can compute $\mathcal{P}(q)$ in $O(\log n)$ time for any query point $q\in\R^d$.\footnote{There is no natural definition of $n$ for the abstract problem $\mathcal{P}$. Nevertheless, all the problems we will eventually consider arise from a set of $n$ objects in $\R^d$ and, for all these problems $O(\log n)$ is the optimal worst-case query time in the models of computation we will consider.} Let $D$ be a probability measure over $\R^d$ representing a distribution of queries. We describe a data structure $T=T_{\mathcal{P},D}$, called the \emph{odds-on tree},\footnote{\textbf{odds-on} $\cdot$ \textit{adj.}\ having a better than even chance of success ``the odds-on favourite''} of size $O(n^\epsilon)$ that can be used as a filter that quickly computes $\mathcal{P}(q)$ for some query values in $\R^d$ and relies on $\mathcal{S}$ for the remaining queries. With an odds-on tree, the expected query time for a point drawn according to $D$ is $O(H^*+1)$, where $H^*$ is a lower-bound on the expected cost of any linear decision tree that solves $\mathcal{P}$. For any constant integer $p\ge 1$, a standard lifting technique allows us to lift queries from $\R^d$ into $\R^{d'}$, with $d'={p+d \choose p}-1$ so that there is a correspondence between $d'$-variate linear inequalities (halfspaces in $\R^{d'}$) and $d$-variate polynomial inequalities of maximum degree $p$ \cite{yy85}. This technique yields the same result, except that $H^*$ becomes a lower-bound on the expected cost of any degree $p$ algebraic decision tree that solves $\mathcal{P}$. Odds-on trees have a plethora of applications, including distribution-sensitive data structures for point location in 2-d, point-in-polytope testing in $d$ dimensions, ray shooting in simple polygons, ray shooting in polytopes, nearest-neighbour queries in $\R^d$, point-location in arrangements of hyperplanes in $\R^d$, and many other geometric searching problems that can be solved in the linear-decision tree model. Furthermore, a variant of the odds-on tree that works in the comparison tree model provides distribution-sensitive data structures for orthogonal searching problems in the comparison tree model. The remainder of this paper is organized as follows: \Secref{prelim} presents some preliminary definitions and background material. \Secref{data-structure} presents the odds-on tree and algorithms for constructing it. \Secref{analysis} proves that the odds-on tree matches the query time of any linear decision tree. \Secref{applications} presents some of the geometric applications of this data structure. Finally, \secref{conclusions} summarizes and concludes with directions for future work. \section{Preliminaries} \seclabel{prelim} Throughout this paper, the underlying dimension, $d$, is a constant, and other constants defined in the paper may (implicitly) depend on $d$. A \emph{simplex} in $\R^d$ is the common intersection of a set of at most $d+1$ closed halfspaces in $\R^d$. Note that, under this definition, simplices need not be bounded and $\R^d$, as well as $\emptyset$, are both simplices. Throughout this paper, we assume an underlying probability measure $D$ over $\R^d$. All expectations and probabilities are (implicitly) with respect to $D$. For any subset $X\subseteq\R^d$, $\Pr(X)$ refers to $D(X)$. We use the notation $D_{|X}$ to denote the distribution $D$ conditioned on $X$, i.e., $D_{|X}(Y)=\Pr(Y\mid X)=\Pr(X\cap Y)/\Pr(X)$ for all $Y\subseteq\R^d$. If $F=\{X_1,\ldots,X_k\}$ are non-overlapping subsets of $\R^d$ then the \emph{entropy} of $F$, denoted $H(F)$ is \[ H(F) = \sum_{i=1}^k \Pr(X_i|{\cup F})\log(1/\Pr(X_i|{\cup F})) \enspace , \] where for any set $S$, $\cup S$ denotes $\bigcup_{s\in S} s$. The probability measure $D$ is used as an input to our algorithms. We assume that the algorithm has access to $D$ through a \emph{Sampling Oracle} that allows us to draw a random sample $q\in\R^d$ from the distribution $D$. A \emph{query problem} over $\R^d$ is a function $\mathcal{P}:\R^d\rightarrow A$ where $A$ is some set of \emph{answers}. We assume the existence of two oracles that allow access to $\mathcal{P}$. The \emph{Backup Oracle} allows us to compute $\mathcal{P}(q)$ for any query point $q$, but requires $O(\log n)$ time to do so. The \emph{Interference Oracle} allows us to test, for any simplex $\Delta$, if there exists $p,q\in\Delta$ with $\mathcal{P}(p)\neq\mathcal{P}(q)$. The running time of the Interference Oracle will be unspecified. The oracles are used in the following ways: The Backup Oracle is used to answer any queries that can not be answered directly by the odds-on tree. The Sampling Oracle and the Interference Oracle are used only during the construction of the odds-on tree. The running times in theorems in Sections~\ref{sec:data-structure} and \ref{sec:analysis} all specify the number of invocations of the Sampling and Interference Oracle used. When discussing applications in \secref{applications}, efficient implementations of the Interference Oracle for specific problems will be described. A \emph{decision tree} for $\mathcal{P}$ is a rooted ordered binary tree in which each internal node $v$ is labelled with a function $v_f:\R^d\rightarrow\{0,1\}$ and each leaf $w$ is labelled with an element $w_A\in A$. A query point $q\in\R^d$ follows a root-to-leaf path, proceeding to the right child of $v$ if $v_f(q)=1$ and the left child of $v$ if $v_f(q)=0$. For a decision tree $T$ and a point $q\in\R^d$, we denote by $T(q)$ the label of the leaf on the root-to-leaf path for $p$ in $T$. A decision tree \emph{solves} $\mathcal{P}$ if $T(q)=\mathcal{P}(q)$ for all $q\in\R^d$. The \emph{depth} of a node $v$ in a rooted tree $T$, denoted $\mathrm{depth}_T(v)$, is the number of edges on the path from $v$ to the root of $T$. The \emph{(expected) cost} of a decision tree, denoted $\mu_D(T)$, is the expected depth of the leaf reached when $p$ is drawn according to the probability measure $D$. Decision trees are classified based on the types of functions, $v_f$, used at their nodes. In a \emph{linear decision tree}, each $v_f$ is a linear inequality. In a \emph{degree $p$ algebraic decision tree}, each $v_f$ is a $d$-variate polynomial inequality of degree $p$. In a \emph{comparison tree}, $v_f$ is a simple comparison that compares one coordinate of a query point $q$ to some value. Entropy, query problems, and decision trees are all related by (half of) Shannon's Source Coding Theorem \cite{s48}: \begin{thm}[Shannon 1948]\thmlabel{shannon} Let $\mathcal{P}:\R^d\rightarrow A$ be a query problem and define $\mathcal{P}^{-1}(a)=\{q\in\R^d:\mathcal{P}(q)=a\}$, for any $a\in A$. Then, for any decision tree $T$ that solves $\mathcal{P}$, \[ \mu_D(T) \ge H(\{\mathcal{P}^{-1}(a): a\in A\}) \enspace . \] \end{thm} \section{The Data Structure} \seclabel{data-structure} In this section we describe a data structure $T_{\mathcal{P},D}$ called the \emph{odds-on tree} that, in conjunction with a Backup Oracle, yields a data structure that solves $\mathcal{P}$ and has expected query time that is within a constant factor of the expected cost of any linear decision tree $T^*$ for $\mathcal{P}$. For a reader familiar with Matou\v{s}ek's efficient partition trees \cite{m92}, the executive summary of this section is as follows: An odds-on tree is essentially a partition tree on a sample of $O(n^\epsilon)$ points drawn according to the distribution $D$. A complete description follows. \begin{thm}[Matou\v{s}ek 1992]\thmlabel{point-partition} There exists a constant $c$ such that, for any set $S$ of $m$ points in $\R^d$ and any constant $r\le m$, there exists a sequence $\langle \Delta_1,\ldots,\Delta_r\rangle$ of closed simplices such that $\bigcup_{i=1}^r \Delta_i = \R^d$, \begin{enumerate} \item $|\Delta_i^*\cap S| \le 2m/r$, where $\Delta_i^*=\Delta_i\setminus \left(\bigcup_{j=1}^{i-1}\Delta_j\right)$, and \item For any hyperplane $\ell$, there are at most $cr^{1-1/d}$ elements of $\{\Delta_1,\ldots,\Delta_r\}$ whose interiors intersect $\ell$. \end{enumerate} The sequence of simplices $\Delta_1,\ldots,\Delta_r$ can be computed in $O(m)$ time. \end{thm} Note that Condition~1 of \thmref{point-partition} is not in the original statement of the theorem, but follows from Matou\v{s}ek's incremental construction of $\Delta_1,\ldots,\Delta_r$ \cite{m92}. Let $S$ be a set of $m$ points in $\R^d$. Then the \emph{partition tree} $T_S$ for $S$ is a rooted ordered tree obtained by recursively applying \thmref{point-partition}. The root of $T_S$ has $r$ children corresponding to the simplices $\Delta_1,\ldots,\Delta_{r}$ obtained by applying \thmref{point-partition} to $S$. The $i$th child of the root is itself the root of the partition tree $T_{S\cap\Delta_i^*}$ for $S\cap\Delta_i^*$. This recursive process stops when the set $S$ contains at most 1 point or when the depth exceeds some pre-specified maximum depth $k$. Next we define some regions, $\Delta(v)$, $\includegraphics{poly}(v)$, and $\Xi(v)$, that are associated with each node $v$ of $T_S$. Every node $v$ in $T_S$, except the root of $T_S$, is naturally associated with a simplex $\Delta(v)$ that was obtained from \thmref{point-partition} and that generated $v$. For the root of $T_S$, we define $\Delta(v)=\R^d$. For a node $v$ of $T$ whose ancestors are $v_1,\ldots,v_i$ we define $\includegraphics{poly}(v)=\Delta(v)\cap\bigcap_{j=1}^i \Delta(v_i)$. Note that $\includegraphics{poly}(v)\subseteq\Delta(v)$, and that $\includegraphics{poly}(v)$ is a convex polytope that has $O(((d+1)(i+1))^{\floor{d/2}})=O(i^{\floor{d/2}})$ vertices since it is the intersection of at most $(d+1)(i+1)$ halfspaces \cite{m70}. For a point $q\in\R^d$, the \emph{search path} for $q$ in $T_S$ starts at the root and proceeds to the first child $i$ such that $q\in\Delta_i$ (note that this implies $q\in\Delta_i^*$) and this process is applied recursively until reaching a leaf of $T_S$. In this way, for every node $v$ of the partition tree there is a maximal subset $\Xi(v)\subseteq \R^d$ such that the search path for every point $q\in\Xi(v)$ contains $v$. Note that $\Xi(v)\subseteq \includegraphics{poly}(v)\subseteq \Delta(v)$, but that $\Xi(v)$ is not necessarily convex or even connected. We extend the definitions of $\Xi$, $\includegraphics{poly}$, and $\Delta$ to sets of nodes in $T$ in the natural way; if $V$ is a set of nodes in $T$, then $\Xi(V)=\{\Xi(v):v\in V\}$, $\includegraphics{poly}(V)=\{\includegraphics{poly}(v):v\in V\}$, and $\Delta(V)=\{\Delta(v):v\in V\}$. The following theorem summarizes the properties of the partition tree $T_{S}$ \cite{m92} (each property is inherited from the corresponding property of the simplicial partition in \thmref{point-partition}): \begin{thm}\thmlabel{point-partition-tree} Let $S$ be a set of $m$ points in $\R^d$, let $T_S$ denote the partition tree described above, and let $V_i$ denote the set of at most $r^i$ nodes of $T_S$ at depth $i$. There exists a constant $c$, independent of $r$ and $m$, such that the partition tree $T_S$ has the following properties, for every $i\in\Z$: \begin{enumerate} \item For every node $v\in V_i$, $|S\cap\Xi(v)| \le m(2/r)^i$, and \item For any hyperplane $\ell$, the number of elements in $\includegraphics{poly}(V_i)$ whose interiors intersect $\ell$ is at most $(cr^{1-1/d})^i$. \end{enumerate} The partition tree $T_S$ can be constructed from $S$ in $O(m\log m)$ time. \end{thm} The following is a sampling version of \thmref{point-partition-tree} that we will use in the construction of an odds-on tree: \begin{thm}\thmlabel{prob-partition-tree} Let $S$ be a sample of $m$ points in $\R^d$ i.i.d.\ according to $D$, let $T_D=T_S$ denote the partition tree given by \thmref{point-partition-tree}, and let $V_i$ denote the set of at most $r^i$ nodes of $T_D$ at depth $i$. There exists a constant $c$, independent of $r$, $m$ and $D$, such that with probability at least $1-O(e^{-m^{1/2}})$, $T_D$ has the following properties, for every $i\in\{0,\ldots,\lfloor (1/4)\log_rm\rfloor\}$: \begin{enumerate} \item For every node $v\in V_i$, $\Pr(\Xi(v)) \le (3/r)^i$, and \item For any hyperplane $\ell$, the number of elements in $\includegraphics{poly}(V_i)$ whose interiors intersect $\ell$ is at most $(cr^{1-1/d})^i$. \end{enumerate} The sample partition tree $T_D$ can be constructed in $O(m\log m)$ time plus the cost of $O(m)$ calls to the Sampling Oracle. \end{thm} \def\buildrel {\rm def} \over ={\buildrel {\rm def} \over =} \def\Pr{\Pr} \begin{proof} Condition~2 follows with certainty from \thmref{point-partition-tree}. In this proof we bound the probability of failure for Condition~1. Let $k=\lfloor (1/4)\log_rm\rfloor$. We will use $D_m(A)$ to denote the empirical measure of a set $A\subseteq\R^d$: \[ D_m (A) \buildrel {\rm def} \over = {{|S\cap A|} \over m} \enspace . \] From \thmref{point-partition-tree} we have \[ \sup_{v\in V_i} D_m(\Xi(v)) \le (2/r)^i \enspace . \] Now, \begin{eqnarray*} \Pr \left( \sup_{v\in V_i} D ( \Xi(v) ) > \left(\frac{3}{r}\right)^i \right) &=& \Pr \left( \cup_{v\in V_i} \left[ D ( \Xi(v) ) - D_m(\Xi(v)) > \left(\frac{3}{r}\right)^i - D_m(\Xi(v)) \right] \right) \cr &\le& \Pr \left( \cup_{v\in V_i} \left[ D ( \Xi(v) ) - D_m(\Xi(v)) > \left(\frac{3}{r}\right)^i - \left(\frac{2}{r}\right)^i \right] \right) \cr &\le& \Pr \left( \sup_{v\in V_i} \left( D ( \Xi(v) ) - D_m(\Xi(v)) \right) > r^{-i} \right) \cr &\le& \Pr \left( \sup_{A \in {\mathcal{A}}} \left( D ( A ) - D_m ( A ) \right) > r^{-i} \right) \cr \end{eqnarray*} where $\mathcal{A}$ are sets formed by taking the intersection of $k$ closed simplices and subtracting $k(r-1)$ (possibly empty) simplices from this intersection. This is because $\Xi(V)\subseteq \mathcal{A}$. We can actually handle all levels $i\in \{0,\ldots,k\}$ of the tree at once with the inequality \begin{eqnarray*} \Pr \left( \cup_{i\in\{1,\ldots,k\}}\left[\sup_{v\in V_i} D ( \Xi(v) ) > \left(\frac{3}{r}\right)^i \right] \right ) \leq \Pr \left( \sup_{A \in {\mathcal{A}}} \left( D ( A ) - D_m ( A ) \right) > r^{-k} \right) \enspace . \end{eqnarray*} The class $\mathcal{A}$ for $k=1$ and $r=1$ is the class of all simplices in $\R^d$. Since a simplex in $\R^d$ is the common intersection of $d+1$ halfspaces in $\R^d$, it helps to first consider halfspaces. The set of halfspaces in $\R^d$ has Vapnik-Chervonenkis dimension $d+1$ (this follows from Radon's Theorem \cite{e93}). By Sauer's lemma \cite{s72}\cite[pages~28--29]{dl01}, the number of subsets of an $m$-point set that can be obtained by intersections with halfspaces does not exceed $(m+1)^{d+1}$. A simple combinatorial argument then implies that the number of subsets of an $m$-point set that can be obtained by intersections with simplices does not exceed $(m+1)^{(d+1)^2}$. Assume now general $r$ and $k$. Then the number of subsets of an $m$-point set that can be obtained by intersections with sets from $\mathcal{A}$ does not exceed $(m+1)^{rk(d+1)^2}$, by the same combinatorial argument. By a version of the Vapnik-Chervonenkis inequality \cite{vc71} shown by Devroye \cite{d82}, \[ \Pr \left( \sup_{A \in {\mathcal{A}}} \left| D ( A ) - D_m ( A ) \right| \ge t \right) \le 4 \left( m^2 + 1 \right)^{rk(d+1)^2} \exp\left(4t+4t^2-2mt^2\right) \enspace , \] for all $t>0$. Thus, \begin{eqnarray*} \Pr \left( \cup_{i\in\{1,\ldots,k\}}\left[\sup_{v\in V_i} D ( \Xi(v) ) > \left(\frac{3}{r}\right)^i \right] \right) &\le& 4 \left( m^2 + 1 \right)^{rk(d+1)^2} \exp\left(4r^{-k}+4r^{-2k}-2mr^{-2k}\right) \\ &\le& \exp\left(2+4r^{-k}+4r^{-2k}+rk(d+1)^2\ln(m^2+1)-2mr^{-2k}\right) \\ &\le& \exp\left(10+r(d+1)^2\log_r(m)\ln(m^2+1)-2m^{1/2}\right) ~,\\ & = & O\left(e^{-m^{1/2}}\right) \end{eqnarray*} since $k= (1/4)\log_rm$. \end{proof} Note that, so far, we have not considered the query problem $\mathcal{P}$ at all; the sample partition tree $T_D$ of \thmref{prob-partition-tree} is defined completely in terms of the probability measure $D$. The \emph{odds-on tree} $T_{\mathcal{P},D}$ for $(\mathcal{P},D)$ is obtained in the following way: We start by constructing a sample partition tree $T_D$ as described in \thmref{prob-partition-tree} using the value $m=n^{\tau}$ for some parameter $\tau > 0$ and setting the maximum depth to $k=\lfloor{(1/4)\log_{r} m}\rfloor$. Next, we trim some nodes of $T_{\mathcal{P},D}$. We use the Interference Oracle to test, for each node $v$ of $T_D$, if $\mathcal{P}(p)=\mathcal{P}(q)$ for all pairs of points $p,q\in \includegraphics{poly}(v)$. If so, we remove all the subtrees rooted at the children of $v$, we call $v$ a \emph{terminal leaf}, and we label $v$ with the label $\ell(v)=\mathcal{P}(q)$. Note that the Interference Oracle works for simplices, but $\includegraphics{poly}(v)$ is a polytope. Thus, this test requires decomposing $\includegraphics{poly}(v)$ into simplices and using the Interference Oracle on each simplex. Using the \emph{bottom vertex triangulation} \cite{c88} for this decomposition allows us to decompose $\includegraphics{poly}(v)$ into $O((\mathrm{depth}_T(v))^{\floor{d/2}})=\log^{O(1)} n$ simplices in $O(\log^{O(1)} n)$ time. This yields the following lemma: \begin{lem}\lemlabel{odds-on-tree-construction} For any $n>0$ and any $\tau > 0$, an odds-on tree of size $m=n^\tau$, with maximum depth $k=\floor{(1/4)\log_{r/3} m}$, and having the properties of \thmref{prob-partition-tree} can be constructed in $O(n^\tau\log^{O(1)} n)$ time plus the cost of $O(n^\tau)$ calls to the Sampling Oracle and $O(n^\tau\log^{O(1)} n)$ calls to the Interference Oracle. \end{lem} Using an odds-on tree to answer a query $q\in \R^d$, is easy: We follow the search path for $q$ in $T_{\mathcal{P},D}$ until we either reach a terminal leaf $v$, in which case we output $\ell(v)$, or we reach a non-terminal leaf $w$ after $O(\log n)$ steps, in which case we rely on the Backup Oracle to report $\mathcal{P}(v)$ in $O(\log n)$ time. The correctness of this procedure follows immediately from the definition of terminal and non-terminal nodes. In the next section, we analyze the performance of odds-on trees. \section{Analysis} \seclabel{analysis} In this section, our goal is to lower-bound the expected cost of any linear decision tree that solves $\mathcal{P}$ in terms of the odds-on tree $T_{\mathcal{P},D}$. We accomplish this by decomposing the nodes of $T_{\mathcal{P},D}$ into subsets with some helpful combinatorial properties. An \emph{$i$-set} of a rooted tree $T$ is a set of vertices in $T$ all of which are at depth at most $i$ and in which no vertex in the set is the ancestor of any other vertex in the set. We say that a set of regions $F=\{X_1,\ldots,X_t\}$, $X_i\subseteq\R^d$, is in \emph{$k$-general position} if there is no hyperplane that intersects $k$ or more elements of $F$. \begin{lem}\lemlabel{independent} Let $T_{\mathcal{P},D}$ be the odds-on tree defined in \secref{data-structure}, let $V$ be an $i$-set of $T_{\mathcal{P},D}$, and let $k>1$ be a constant. Then $V$ contains a subset $V'\subseteq V$ such that the elements of $\includegraphics{poly}(V')$ are pairwise disjoint and in $k$-general position and $|V'| = \Omega(|V|/r^{i(d/k+1-1/d+\delta)})$, where $\delta > 0$ is a decreasing function of $r$. \end{lem} \begin{proof} We will first use the probabilistic method \cite{as08} to establish the existence of a (not necessarily disjoint) set $V''$ satisfying the size and $k$-general position requirements and then show that $V''$ contains a large subset $V'$ whose elements are also pairwise disjoint. Let $V''$ be a Bernoulli sample of $V$ where each element is selected independently with probability $p=r^{-i(d/k+1-1/d+\delta)}$. We will prove that \[ \Pr\left\{ \mbox{$\includegraphics{poly}(V'')$ is in $k$-general position and $|V''| = \Omega(p|V|)$} \right\} > 0 \enspace . \] Consider any hyperplane $\ell$. Condition~2 of \thmref{prob-partition-tree} implies that $\ell$ intersects the interior of at most $(cr^{1-1/d})^{i}$ elements of $V$ for some constant $c$. The probability that $\ell$ intersects the interior of $k$ or more elements of $V''$ is therefore no more than \[ \binom{((cr)^{1-1/d})^{i}}{k}\cdot p^k \le (cr)^{i(k-k/d)}p^k \enspace . \] For each node $v\in V$, $\includegraphics{poly}(v)$ has $O(i^{\floor{d/2}})$ vertices, and the number of nodes in $V$ is at most $r^{i}$. Therefore, the elements of $\includegraphics{poly}(V)$ define a \emph{test set} $L$ of $O((i^{\floor{d/2}}r^i)^d)=O(i^{d^2/2}r^{di})$ hyperplanes such that $\includegraphics{poly}(V'')$ is in $k$-general position if and only if no hyperplane in $L$ intersects $k$ or more elements of $\includegraphics{poly}(V'')$. The probability that \emph{any} hyperplane in $L$ intersects $k$ or more elements of $\includegraphics{poly}(V'')$ is therefore at most \[ O(i^{d^2/2}r^{di})\cdot (cr)^{i(k-k/d)}p^k = O(r^{i(d+k-k/d+\delta)})p^k = O(r^{\delta-k\delta)}) = o(1) \enspace \] for any $\delta > (k-k/d)\log_r c$. The above argument shows that the nodes in $V''$ are quite likely to be in $k$-general position. To see that $V''$ is sufficiently large, we simply observe that $|V''|$ is a $\mathrm{binomal}(|V|,p)$ random variable and therefore has median value at least $\lfloor{p|V|}\rfloor = \Omega(|V|/r^{i(d/k+1-1/d+\delta)})$. Therefore, \[ \Pr\left\{ \mbox{$\includegraphics{poly}(V'')$ is in $k$-general position and $|V''| = \Omega(|V|/r^{i(d/k+1-1/d+\delta)})$ } \right\} \ge 1- (o(1) + 1/2) > 0 \enspace . \] This establishes the existence of a sufficiently large set $V''$ such that $\includegraphics{poly}(V'')$ is in $k$-general position. Finally, we select $V'\subseteq V''$ so that the elements of $\includegraphics{poly}(V')$ are pairwise disjoint. To do this, imagine sweeping a hyperplane $\ell(t)=\{(x_0,\ldots,x_d)\in\R^d:x_0=t\}$ from $t=-\infty$ to $+\infty$. Associate with each element $v\in V'$, the maximal interval $[a_v,b_v]$ such that $\includegraphics{poly}(v)$ intersects $\ell(t)$ for all $t\in[a_v,b_v]$. This yields a set $S_{V''}$ of real intervals such that no point is contained in $k$ or more elements of $S_{V''}$. Dilworth's Theorem \cite{d50} implies that $S_{V''}$ contains a subset of size at least $|V''|/k$ of non-overlapping intervals. This subset corresponds to a subset $V'\subseteq V''$ with $|V'|\ge |V''|/k$ and such that the elements of $\includegraphics{poly}(V')$ are pairwise disjoint. The set $V'$ satisfies all the conditions of the lemma. \end{proof} We are now ready to show that the expected search time in the odds-on tree is a lower bound on the expected cost of any linear decision tree that solves $\mathcal{P}$. \begin{lem}\lemlabel{lower-bound} Let $T_{\mathcal{P},D}$ be the odds-on tree defined in \secref{data-structure} and assume $T_{\mathcal{P},D}$ satisfies Conditions~1 and 2 of \thmref{prob-partition-tree}. Let $L$ denote the set of leaves of $T_{\mathcal{P},D}$, and let $T^*$ be any linear decision tree that solves $\mathcal{P}$. Then \[ \mu_D(T^*) = \Omega(H(\Xi(L))-1) \enspace . \] \end{lem} \begin{proof} This proof mixes the ideas from the proofs of Lemma~3 by Dujmovi\'c \etal\ \cite{dhm09} and Lemma~4 by Collette \etal\ \cite{cdilm08}. For an $i$-set $V$ of nodes in $T_{\mathcal{P},D}$, we define the shorthands $\Pr(V)=\Pr(\cup\Xi(V))$ and $H(V)=H(\Xi(V))$. Let $T'$ be the tree obtained from $T_{\mathcal{P},D}$ by removing all terminal leaves, and let $L'$ denote the set of leaves of $T'$. Note that \[ H(L') = H(L) - O(\log r) = H(L) - O(1) \] since each leaf in $L'$ has at most $r=O(1)$ children in $L$. We first partition $L'$ into groups $G_1,G_2,\ldots$, where $G_i$ contains all leaves $v$ such that $1/2^{i-1}\ge \Pr(\Xi(v)) \ge 1/2^{i}$. Note that Condition~1 of \thmref{prob-partition-tree} implies that the depth of a node in $G_i$ is at most $i/\log(r/3)$. We then further partition each group $G_i$ into subgroups $G_{i,1},\ldots,G_{i,t_i}$. For each $j\in\{0,\ldots,G_{i,t_i-1}\}$, $|G_{i,j}|= \Omega(2^{\alpha i})$, for some constant $\alpha >0$ and the elements of $\includegraphics{poly}(G_{i,j})$ are pairwise disjoint and in $k$-general position. Furthermore, the final subgroup, $G_{i,t_i}$ has size at most $O(2^{\beta i})$, for some constant $\beta < 1$. This partitioning is accomplished by repeatedly applying \lemref{independent} to remove a subset $G_{i,j}\subseteq G_{i}$ that is in $k$-general position and has size $\Omega(2^{\alpha i})$, stopping the process once the size of $G_i$ drops below $O(2^{\beta i})$. This works provided that we choose $\beta$, $k$, and $r$ so that $\beta > ((\log r)/(\log (r/3) ))((d/k+1-1/d+\delta)$ and set $\alpha\le\beta - ((\log r)/(\log (r/3)))(d/k+1-1/d+\delta)$. For example, by choosing $r$ and $k$ to be sufficiently large, $\beta=1-1/(3d)$ and $\alpha=1/(3d)$. Consider the linear decision tree $T^*$ that solves $\mathcal{P}$. The leaves of $T^*$ partition $\R^d$ into cells whose closures are convex polytopes. For a leaf $w$ of $T^*$ we denote its polytope by $\includegraphics{poly}(w)$. If the depth of $w$ is $i$, then $\includegraphics{poly}(w)$ is the intersection of at most $i$ halfspaces. This implies that $\includegraphics{poly}(w)$ intersects at most $ik$ elements of $\includegraphics{poly}(G_{i,j})$ since, otherwise, $\includegraphics{poly}(w)$ contains $\includegraphics{poly}(v)$ for some non-terminal node $v\in G_{i,j}$. (This would contradict the assumption that $T^*$ solves $\mathcal{P}$.) Let $w$ be some leaf of $T^*$ such that $\includegraphics{poly}(w)$ intersects $t$ elements of $\includegraphics{poly}(G_{i,j})$. Then, by the discussion in the previous paragraph, $\mathrm{depth}_{T^*}(w) \ge \ceil{t/k}$. We can easily create a subtree $T_w^*$ of $w$ whose height is at most $t$ and with the property that $\includegraphics{poly}(w')$ intersects at most one element of $\includegraphics{poly}(G_{i,j})$ for each leaf $w'$ of $T_w^*$.\footnote{For example, making the root of $T^*_w$ correspond to the bisector of two polyhedra of $\includegraphics{poly}(G_{i,j})$ that intersect $\includegraphics{poly}(w)$ means that each of the children of the root intersect at most $t-1$ elements of $\includegraphics{poly}(G_{i,j})$. Applying this recursively yields a subtree of height at most $t-1$. A slightly more involved argument can produce a tree $T^*_w$ of depth $O(\log t)$.} If we do this for every leaf $w$ of $T^*$ we obtain a tree $T^{*}_{i,j}$ such that every leaf of $T^*_{i,j}$ intersects at most one element of $\includegraphics{poly}(G_{i,j})$. Let $D_{i,j}=D_{|\cup\Xi(G_{i,j})}$ denote the distribution $D$ conditioned on $\cup\Xi(G_{i,j})$. Note that the leaves of $T^*_{i,j}$ could be relabeled so that they indicate which element of $\includegraphics{poly}(G_{i,j})$ (if any) that they intersect. By Shannon's Theorem (\thmref{shannon}), this implies that \[ (k+1)\cdot\mu_{D_{i,j}}(T^*) \ge \mu_{D_{i,j}}(T^*_{i,j}) \ge H(G_{i,j}) \enspace . \] It follows \cite[Lemma~3]{cdilm09} that \[ (k+1)\cdot\mu_D(T^*) \ge H(L') - H(\{\cup \Xi(G_{i,j}):i\in\N,\, j \in\{1,\ldots,t_{i}\}) - O(1) \enspace . \] Thus, all that remains is to upper-bound the contribution of $\bar{H}=H(\{\cup \Xi(G_{i,j}):i\in\N,\, j \in\{1,\ldots,t_{i}\})$, as follows: \begin{eqnarray*} \bar{H} &= & H(\{\cup G_{i,j}:i\in\N,\, j \in\{1,\ldots,t_{i}\}) \\ & = & \sum_{i=1}^\infty \sum_{j=1}^{t_{i}} \Pr(G_{i,j})\log(1/\Pr(G_{i,j})) \\ & = & \sum_{i=1}^\infty \left( \sum_{j=1}^{t_{i}-1} \Pr(G_{i,j})\log(1/\Pr(G_{i,j}) + \Pr(G_{i,t_i})\log(1/\Pr(G_{i,t_i})) \right) \\ & \le & \sum_{i=1}^\infty \left( \sum_{j=1}^{t_{i}-1} \Pr(G_{i,j})\log(2^{i-\alpha i}) + i 2^{\beta i - i + 1} + O(1) \right) \\ & \le & (1-\alpha)H(L') + O(1) \enspace . \end{eqnarray*} Thus, we have \begin{eqnarray*} (k+1)\mu_D(T^*) &\ge& H(L') - \bar{H} -O(1) \\ &\ge& \alpha H(L') - O(1) \\ &\ge& \alpha H(L) - O(1) \\ & = & \Omega(H(L) - 1) \enspace , \end{eqnarray*} so $\mu_D(T^*) = \Omega(H(L) - 1)$, as required. \end{proof} \noindent\textbf{Remark:} By more carefully handling the constants in the proof of \lemref{lower-bound}, and allowing $k$ and $r$ to be arbitrarily large, one can obtain the tighter lower-bound \[ \mu_D(T^*) + \log (\mu_D(T^*)) \ge \alpha H - O(\log (kr)) \] where $\alpha$ can be made arbitrarily close to $1/d$ by increasing $k$ and $r$. \begin{thm}\thmlabel{odds-on} Let $\mathcal{P}:\R^d\rightarrow A$ be a decision problem for which we have a ($O(\log n)$ time) Backup Oracle and an Interference Oracle, and let $D$ be any probability measure over $\R^d$ for which we have a Sampling Oracle. Then, for any constant $\epsilon > 0$, an odds-on tree of size $O(n^\epsilon)$ can be constructed in $O(n^\epsilon)$ time plus the cost of $O(n^\epsilon)$ calls to the Sampling Oracle and $O(n^\epsilon)$ calls to the Interference Oracle. This odds-on tree can, in conjunction with the Backup Oracle, compute $\mathcal{P}(q)$ for any $q\in\R^d$ drawn according to $D$ in $O(H^*+1)$ expected time, where $H^* \le \mu_D(T^*)$ for any linear decision tree $T^*$ that solves $\mathcal{P}$. \end{thm} \begin{proof} Applying \lemref{odds-on-tree-construction} with $\tau < \epsilon$ yields the stated bounds on the construction time and the use of the Sampling and Interference Oracles. If $T_{\mathcal{P},D}$ satisfies Conditions~1 and 2 of \thmref{prob-partition-tree} then, by \lemref{lower-bound}, the expected time to answer queries using $T_{\mathcal{P},D}$ is \[ \sum_{t\in L} \Pr(t)O(\mathrm{depth}_T(t)) = \sum_{t\in L}\Pr(t)O(\log(1/\Pr(t))) = O(H(L)) \enspace . \] Otherwise, the expected time to answer queries using $T_{\mathcal{P},D}$ is $O(\log n)$. Therefore, the expected time (where the expectation is taken over $D$ and the random sampling used to generate $S$ in \thmref{prob-partition-tree}) to answer a query using an odds-on tree is \[ (1-O(e^{-n^{\tau/2}}))\cdot O(H(L)) + O(e^{-n^{\tau/2}})\cdot O(\log n) = O(H(L)+1) \enspace . \] On the other hand, by \lemref{lower-bound} the expected cost of any linear decision tree $T^*$ that solves $\mathcal{P}$ is \[ \mu_D(T^*) = \Omega(H(L) - 1) \enspace , \] which completes the proof. \end{proof} Using a standard lifting of query points \cite{yy85}, any degree $p$ algebraic decision tree that solves a problem $\mathcal{P}:\R^d\rightarrow A$ can be implemented as a linear decision tree in $\R^{d'}$, with $d'={p+d\choose p}-1$. Applying this yields the following corollary: \begin{cor}\corlabel{odds-on-algebraic} Let $\mathcal{P}:\R^d\rightarrow A$ be a decision problem for which we have an ($O(\log n)$ time) Backup Oracle and an Interference Oracle, and let $D$ be any probability measure over $\R^d$ for which we have a Sampling Oracle. Then, for any constants $\epsilon > 0$ and $p\ge 1$, using $O(n^\epsilon)$ calls to the Sampling Oracle and $O(n^\epsilon)$ calls to the Interference Oracle, an odds-on tree can be constructed in $O(n^\epsilon)$ time and $O(n^\epsilon)$ space. This odds-on tree can, in conjunction with the backup oracle, answer $\mathcal{P}$-queries drawn according to $D$ in $O(H^*+1)$ expected time, where $H^* \le \mu_D(T^*)$ for any degree $p$ algebraic decision tree $T^*$ that solves $\mathcal{P}$. \end{cor} \section{Applications} \seclabel{applications} In this section, we discuss a few of the many potential applications of odds-on trees. In all of our applications, the problem $\mathcal{P}$ is a query problem over some set of $n$ geometric objects in $\R^d$. In order to shorten the statements of the theorems in this section, we say that a data structure for a problem $\mathcal{P}$ is \emph{distribution-sensitive (in the linear decision tree model)} if the expected query time of the data structure is $O(\mu_D(T^*)+1)$ for any linear decision tree $T^*$ that solves $\mathcal{P}$. Before delving into the details of the applications, we first outline some general strategies for implementing the Interference Oracle needed to build an odds-on tree. An odds-on tree can be made to have size $O(n^\epsilon)$ for any constant $\epsilon > 0$. Furthermore, the number of calls to the Interference Oracle made during the construction of an odds-on tree is $O(n^\epsilon)$. In almost all of our applications, the Interference Oracle can be trivially implemented to run in $O(n)$ time by testing the query simplex $\Delta$ against each input element. This means that, even with no preprocessing, the contribution of calls to the Interference Oracle to the construction time of an odds-on tree is no more than $O(n^{1+\epsilon})$. Some of the subsequent theorems in this paper will use this fact implicitly. In other cases, the Interference Oracle corresponds to a natural query for which there exists (or we can develop) an $O(n^{1-\epsilon})$ query-time $O(n\log n)$ preprocessing-time data structure. In these cases, the time to construct the odds-on tree becomes $O(n\log n)$. One particularly common instance of this occurs when Interference Oracle queries can be reduced to $O(1)$ simplex range counting queries in $\R^{d'}$ for some constant dimension $d'$. In this case, Matou\v{s}ek's partition trees \cite{m92} yield an Interference Oracle that can be constructed in $O(n\log n)$ time and that can answer queries in $O(n^{1-1/d+\epsilon})$ time. \subsection{Point Location Problems} The \emph{planar point location problem} is to determine which face of a planar straight line graph $G$ contains a query point $q\in\R^2$. A number of authors have considered distribution-sensitive algorithms for this problem \cite{acmr00,amm00,amm01a,amm01b,ammw07,cdilm08,i01,i04} and have mostly solved it. The most general such result is due to Collette \etal\ \cite{cdilm08} and gives a distribution-sensitive data structure for point location in connected planar subdivisions. This leaves open the case where the graph, $G$, of the subdivision is not connected. Since there are several $O(\log n)$ query-time, $O(n\log n)$ preprocessing-time data structures for planar point location in (possibly disconnected) planar subdivision \cite{as98,egs86,k83,m90,st86} we can apply odds-on trees. To obtain a fast preprocessing time, we can use an implementation of the Interference Oracle based on partition trees. For this problem, the Interference Oracle must answer queries of the form: ``Does the interior of query triangle $\Delta$ intersect more than one face of $G$?'' We can build a data structure for this problem by first removing from $G$ any edges and vertices not on the boundary of more than 1 face to obtain a graph $G'$. The interior of $\Delta$ intersects more than one face of $G$ if and only if the interior of $\Delta$ intersects an edge of $G'$. If $\Delta$ intersects an edge of $G'$ then $\Delta$ contains a vertex of $G'$ or some edge of $\Delta$ intersects some edge of $G'$. Determining if $\Delta$ contains a vertex of $G'$ is the classic 2-dimensional simplex-range counting problem that can be solved in $O(n^{1/2+\epsilon})$ time after $O(n\log n)$ preprocessing using partition trees. To determine if some edge $uw$ of $\Delta$ intersects some edge $xy$ of $G'$, we observe that $uw$ and $xy$ intersect if and only if \[ L(u,w,y) \wedge L(x,y,u) \wedge L(w,u,x) \wedge L(y,x,w) \] or \[ R(u,w,y) \wedge R(x,y,u) \wedge R(w,u,x) \wedge R(y,x,w) \] where $L(a,b,c)$ and $R(a,b,c)$, are predicates that are true if and only if the sequence of points $abc$ form a left turn, respectively, a right turn. If we fix $x$ and $y$, then each of the above predicates is the sign of a linear function over the variables $u_1$, $u_2$, $w_1$, and $w_2$. Therefore, we can treat each edge of $G'$ as a point in $\R^4$, yielding a point set $S_{G'} \subset\R^4$ such that testing if edge $uw$ intersects some edge of $G'$ can be reduced to two simplex range counting queries over $S_{G'}$. Again, partition trees allow us to do this in $O(n^{3/4+\epsilon})$ time after $O(n\log n)$ preprocessing. Therefore, the Interference Oracle for point location can be implemented to run in $O(n^{3/4+\epsilon})$ time after $O(n\log n)$ preprocessing. This yields our first theorem: \begin{thm}\thmlabel{planar-point-location}\thmlabel{first-app}\thmlabel{a} There exists a distribution-sensitive data structure for the planar point location problem that uses $O(n\log n)$ preprocessing time and $O(n)$ space. \end{thm} The \emph{3-d point in polytope problem} is the problem of determining if a query point $q\in\R^3$ is contained in a 3-dimensional polytope $P$. The Dobkin-Kirkpatrick hierarchy \cite{dk83} gives an $O(\log n)$ query-time, $O(n)$ preprocessing-time data structure for this problem. Furthermore, Interference Oracle queries (which involve testing if a tetrahedron intersects the boundary of $P$) can also be answered in $O(\log n)$ time by the Dobkin-Kirkpatrick hierarchy. \begin{thm}\thmlabel{b} There exists a distribution-sensitive data structure for the 3-dimensional point in polytope problem that uses $O(n)$ preprocessing time and space. \end{thm} These results can be extended to higher dimensions, though with more space \cite{c88}: \begin{thm} There exists a distribution-sensitive data structure for the $d$-dimensional point in polytope problem that uses $O(n^{\floor{d/2}+\epsilon})$ preprocessing time and space. \end{thm} The problem of \emph{point-location in an arrangement of hyperplanes} is the problem of determining which cell in an arrangement of $n$ hyperplanes in $\R^d$ contains a query point $q\in\R^d$. Liu gives an $O(n^d)$ space and preprocessing-time, $O(\log n)$ query-time data structure for this problem \cite{l04}. \begin{thm} There exists a distribution-sensitive data structure for point location in an arrangement of hyperplanes that uses $O(n^d)$ preprocessing time and space. \end{thm} \subsection{Post-Office Queries} For an $n$ point set $S\subset\R^d$, the \emph{$d$-dimensional post-office problem} asks for a point of $p\in S$ that minimizes the Euclidean distance $\|pq\|$ for a query point $q\in\R^d$. The post-office problem can be solved efficiently through the use of point location in Voronoi diagrams. In 2-dimensions, Voronoi diagrams are planar graphs of size $O(n)$ and can be computed in $O(n\log n)$ time \cite{ps85}. Combining this with \thmref{planar-point-location} gives a distribution-sensitive data structure for the 2-d post-office problem: \begin{thm}\thmlabel{c} There exists a distribution-sensitive data structure for the 2-dimensional post office problem that uses $O(n\log n)$ preprocessing time and $O(n)$ space. \end{thm} In $d >2$ dimensions, the post-office problem can still be solved using Voronoi diagrams, but the space and preprocessing costs are higher \cite{c88postoffice}: \begin{thm} There exists a distribution-sensitive data structure for the 2-dimensional post office problem that uses $O(n^{\ceil{d/2}+\epsilon})$ preprocessing time and space. \end{thm} \subsection{Ray Shooting} The \emph{ray shooting in a polygon problem} asks for the first point on the boundary of a polygon $P$ intersected by a query ray. A query ray can be represented as a pair of points in $\R^2$ (the source and any other point on the ray), so this is a query problem over $\R^4$. Several $O(n)$ preprocessing time $O(\log n)$ query time solutions to this problem exist \cite{cegghss94,hs95}. \begin{thm} There exists a distribution-sensitive data structure for the ray shooting in a polygon problem that uses $O(n^{1+\epsilon})$ preprocessing time and $O(n)$ space. \end{thm} The \emph{ray shooting in a $d$-dimensional polytope problem} asks for the first point on the boundary of a convex polytope $P\subseteq\R^d$ intersected by a query ray. For polytopes in $\R^3$, the Dobkin-Kirkpatrick hierarchy \cite{dk83} can perform ray shooting in $O(\log n)$ time per query. \begin{thm} There exists a distribution-sensitive data structure for the ray shooting in a $3$-dimensional polytope problem that uses $O(n^{1+\epsilon})$ preprocessing time and $O(n)$ space. \end{thm} Schwarzkopf \cite{s92} gives an $O(n^{\floor{d/2}+\epsilon})$ space, $O(\log n)$ query time solution for the problem of ray shooting in a $d$-dimensional polytope with $d \ge 4$. \begin{thm} There exists a distribution-sensitive data structure for the ray shooting in a $d$-dimensional polytope problem that uses $O(n^{\floor{d/2}+\epsilon})$ preprocessing time and space. \end{thm} \subsection{Orthogonal Range Counting} The \emph{2-d orthogonal range counting problem} is the problem of counting the number of points of a data set $S\subset\R^2$ that are contained in a query rectangle $[q_1,q_2]\times[q_3,q_4]$. (Note that this produces a query point $q\in \R^4$.) Bentley's range trees \cite{b75}, with fractional cascading \cite{cg86,l78}, yield an $O(n\log n)$ preprocessing time and space, $O(\log n)$ query time data structure for this problem. Interference Oracle queries for 2-d orthogonal range counting require determining if any rectangle represented by a point in a 4-dimensional simplex has some point of $S$ on its boundary. This problem can be decomposed into $O(1)$ simplex range counting queries in $\R^4$ in a manner similar to that used for the point location problem. Thus, an Interference Oracle for this problem can be implemented in $O(n^{3/4+\epsilon})$ time after $O(n\log n)$ preprocessing. \begin{thm}\thmlabel{d} There exists a distribution-sensitive data structure for the 2-d orthogonal range counting problem that uses $O(n\log n)$ preprocessing time and $O(n\log n)$ space. \end{thm} Range trees satisfy the constraints of the comparison tree model of computation, which is considerably weaker than the linear decision tree model. Dujmovi\'c \etal\ \cite{dhm09} give a distribution-sensitive 2-d orthogonal range counting data structure that works in the comparison tree model. However, their technique can only answer \emph{2-sided queries}, i.e., queries of the form $[q_1,\infty)\times[q_2,\infty)$. Filter trees can be made to work in the comparison tree model and handle full (4-sided) 2-d orthogonal range counting queries. The key modification required is the use of $k$-d trees in \thmref{prob-partition-tree} rather than Matou\v{s}ek's partition trees. (This method, in 2 dimensions, is essentially how Dujmovi\'c \etal\ obtain their results.) In the comparison tree model, an Interference Oracle query is a $d$-dimensional box, rather than a simplex. In the special case of 2-d orthogonal range counting, Interference Oracle queries can be reduced to a constant number of rectangular range counting queries among the input set $S$. Therefore, Interference Oracle queries can, in this case, be answered in $O(\log n)$ time after $O(n\log n)$ preprocessing using range trees. \begin{thm}\thmlabel{last-app} There exists a distribution-sensitive data structure \emph{in the comparison tree model} for the 2-d orthogonal range counting problem that uses $O(n\log n)$ preprocessing time and space. \end{thm} \section{Summary and Conclusions} \seclabel{conclusions} We have presented a data structure --- the odds-on tree --- that can be added on to any data structure that answers queries drawn from some distribution $D$ over $\R^d$. If the underlying data structure has query time $O(\log n)$, then the combined data structure will have optimal expected query time in the linear decision tree model. A variant of the odds-on tree based on $k$-d trees provides a similar result in the comparison tree model. \secref{applications} contains a smattering of applications of the odds-on tree. Many more are possible. Note that \corref{odds-on-algebraic} applies to all of the problems in Theorems~\ref{thm:first-app}--\ref{thm:last-app} to yield, for any constant $p$, data structures that are distribution-sensitive in the degree $p$ algebraic decision tree model. However, when applying \corref{odds-on-algebraic}, the Interference Oracle becomes considerably more complicated, so that the fast preprocessing times in Theorems~\ref{thm:a}, \ref{thm:b}, \ref{thm:c}, and \ref{thm:d} increase to $O(n^{1+\epsilon})$. However, the real power of \corref{odds-on-algebraic} is its applications to problems that cannot be solved by finite linear decision trees. Examples of such problems include fixed-radius circular ray shooting in polygons \cite{cceo04}, point-location in 2-dimensional power diagrams \cite{a87}, point-location in planar subdivisions whose edges are defined by algebraic curves (such as arrangements of circles), and many others. We believe that the odds-on tree may actually be practical in some settings. The comparison-tree version of the odds-on tree is based on $k$-d trees, which are simple and widely used in practice. At worst, a comparison-based odds-on tree will perform $O(\epsilon\log n)$ extra comparisons before falling back to the backup data structure. In low dimensions, simpler alternatives to Matou\v{s}ek's Partition Theorem (\thmref{point-partition}) are available and may be more practical. For example, in 2-d one can, $O(n)$ time, find two lines that partition any $m$ point set into four point sets each of size at most $\ceil{m/4}$ and such that no line intersects more than the 3 of the resulting sets \cite{m85}. This result is strong enough that it can be used in place of \thmref{point-partition} to prove all our results (for 2-d problems) and yields an odds-on tree that only requires 2 point-line comparisons at each node. Because the odds-on tree is so small (of size $O(n^\epsilon)$) it may be useful to speed up searching in environments where memory is constrained. For example, it can be applied to the succinct point location data structures of Bose \etal\ \cite{bchmm09} to obtain a distribution-sensitive and succinct data structure for point location. In external-memory settings, an odds-on tree may also be useful as a filter that is sufficiently small to fit into internal memory while the backup structure lives in (larger but slower) external memory. \bibliographystyle{plain}
\section{Introduction} In this talk, manifolds, maps, group actions {\em e.\ t.\ c.} are all assumed to be of class $C^\infty$. Throughout the talk, $M$ stands for a closed $C^\infty$ manifold, and $G$ for a connected and simply connected Lie group. Denote by ${\mathcal A}^r$ the set of locally free right $G$-actions (of class $C^\infty$) endowed with the Whitney $C^r$-topology ($1\leq r\leq \infty$). An action $\varphi:M\times G \rightarrow M$ in ${\mathcal A}^r$ is said to be $C^r$ {\em locally rigid} if there exists a neighbourhood $\mathcal U$ of $\varphi$ in ${\mathcal A}^r$ such that any $\psi \in \mathcal U$ is smoothly conjugate to $\varphi$ by a diffeomorphism $F$ of $M$, up to an automorphism $A$ of $G$, {\em i.\ e.} \begin{equation} \label{e1} F(\varphi(x;g))=\psi(F(x);A(g)) \end{equation} for any $x\in M$ and $g\in G$. An action $\varphi$ is said to be {\em globally rigid} if (\ref{e1}) holds for any $\psi\in{\mathcal A}^r$. \bigskip Of course this is an extremely strong property. For example any flow (an ${\mathbb R}$-action) on a manifold $M$ other than $S^1$ cannot be $C^1$ locally rigid. This follows from Pugh's closing lemma in case the flow does not admit a periodic orbit. In the other case notice that the eigenvalues of the Poincar\'e map along a periodic orbit can be easily changed by a perturbation. However for other Lie groups, there exist examples of even globally rigid actions. Let $GA$ be the Lie group of the orientation preserving affine transformations on the real line. Let $A\in SL(2;{\mathbb Z})$ be a hyperbolic automorphism of the 2-torus and denote by ${\mathbb T}_A$ the mapping torus of $A$. Then the weak stable foliation of the suspension flow is the orbit foliation of a locally free $GA$ acion. This action is known to be globally rigid (\cite{GS,G}). There are other examples. Let $\Gamma$ be a Fuchsian triangle group. Then on the manifold $\Gamma\setminus PSL(2;{\mathbb R})$, a locally free $GA$-action is defined by the right action of the elements $$ \left[ \begin{array}{cc} e^{t/2} & x \\ 0 & e^{-t/2} \end{array} \right] . $$ The orbit foliation of this action is again a weak stable foliation of an Anosov flow. E. Ghys (\cite{G}) showed that this action is also globally rigid. The proofs of these facts consist of the studies of two independent phenomena. One is concerned about the $C^\infty$ rigidity of the orbit foliation, and the other is about the rigidity of the parametirization of the action. When A. Katok and Lewis (\cite{KL}) showed the $C^1$ local rigidity for certain ${\mathbb R}^n$-actions, the same strategy was taken. So let us devide the local rigidity into two parts. \begin{Definition} {\em An action $\varphi\in{\mathcal A}^r$ is said to be $C^r$ {\em locally orbit rigid} if there exists a neighbourhood $\mathcal U$ of $\varphi$ such that the orbit foliation ${\mathcal O}_\psi$ of any element $\psi\in \mathcal U$ is smoothly conjugate to the orbit foliation ${\mathcal O}_\varphi$ of $\varphi$ {\em i.\ e.} there exists a diffeomorphism $F$ of $M$ which sends each leaf of ${\mathcal O}_\psi$ to a leaf of ${\mathcal O}_\varphi$.} \end{Definition} \begin{Definition} {\em An action $\varphi\in {\mathcal A}^r$ is said to be {\em parameter rigid} if for any action $\psi\in {\mathcal A}^r$ such that ${\mathcal O}_\psi={\mathcal O}_\varphi$, there exist a diffeomorphism $F$ which preserves leaves of these identical foliation and an automorphism $A$ of $G$ such that (\ref{e1}) holds.} \end{Definition} The local version of parameter rigidity is not defined, simply because we have no intermediate example. In Sect.\ 2, 3 and 10, we will focus our attention on the parameter rigidity for abelian or solvable Lie group actions. \bigskip We will define the leafwise cohomology in Sect.\ 3 and discuss its close relation with the parameter rigidity when the group $G$ is abelian. After we introduced methods for the computation of the leafwise cohomology in Sect.\ 4, Sect.\ 5, 6, 8 and 9 are devoted to the computational results for various concrete foliations. In Sect.\ 10, we raise examples of parameter rigid solvable group actions. The final Sect.\ 11 is devoted to the relation of the leafwise cohomology to the problem of the existence of Riemannian metric for which all the leaves are minimal surfaces. \section{Regidity of flows} An ${\mathbb R}$-action $\varphi$ is what is usually called a flow, and is associated with the vector field $X$ given by $$ X_x=\frac{d}{dt}\vert_{t=0}\varphi^t(x). $$ An ${\mathbb R}$-action $\varphi$ is locally free if and only if $X$ is nonsingular. As we have mentioned in Sect.\ 1, there is no $C^1$ locally rigid flow unless the manifold is $S^1$. However there does exist a parameter rigid flow, which we shall explain below. Given a real number $\alpha\in{\mathbb R}$, define a Kroneker flow $\varphi_\alpha$ on the 2-torus $T^2$ by $$ \varphi^t_\alpha(x,y)=(x+\alpha t, y+t). $$ If the slope $\alpha$ is rational, then the flow is periodic {\em i.\ e.} $\varphi^q$ is the identity for some $q>0$. All the orbits are closed and their periods are the same. The flow can never be parameter rigid, since one can change the periods of closed orbits so that they are not identical. If $\alpha$ is irrational, then all the orbits are dense in $T^2$. It is also known that the flow is uniquely ergodic {\em i.\ e.} the $\varphi_\alpha$-invariant probability is unique. As for parameter rigidity, we have the following dichotomy. \begin{Proposition} \label{p1} The Kronecker flow $\varphi_\alpha$ is parameter rigid if and only if the slope $\alpha$ is badly approximable. \end{Proposition} \begin{Definition} {\em A real number $\alpha$ is said to be {\em badly approximable} if there exist $C>0$ and $\rho>0$ such that $$ \Vert k\alpha\Vert_{S^1}>C\vert k\vert^{-\rho}, \ \ \ \forall k \in {\mathbb Z}\setminus \{0\}, $$ where $\Vert \cdot \Vert_{S^1}$ denotes the distance to 0 of the projected image in $S^1={\mathbb R}/{\mathbb Z}$.} \end{Definition} The proof of Proposition \ref{p1} in one direction is in order. Assume $\alpha$ is badly approximable. Let $S^1$ be a circle in $T^2$ defined by $y=0$. $S^1$ is a global cross section for $\varphi_\alpha$ and the first return map is the rotation by $\alpha$, $R_\alpha$. Let $\psi$ be the flow obtained by reparametrization of $\varphi_\alpha$, {\em i.\ e.} ${\mathcal O} _\psi={\mathcal O}_{\varphi_\alpha}$. Let $f:S^1\rightarrow R$ be the first return time of $\psi$ for the cross section $S^1$; thus $\psi^{f(x)}(x)=R_\alpha(x)$. \medskip \noindent \textsc{Claim:} {\em There exist $g\in C^\infty(S^1)$ and $c\in{\mathbb R}$ such that $f=g\circ R_\alpha-g+c$.} \medskip Let us show first that Claim is sufficient to show the parameter rigidity of $\varphi_\alpha$. For this, define a global cross section $C$ to the flow $\psi$ by $$ C=\{ \psi^{-g(x)}(x)\mid x\in S^1\}. $$ Then we have $$ \psi^c(\psi^{-g(x)}(x))=\psi^{-g(R_\alpha(x))}(\psi^{f(x)}(x)) =\psi^{-g(R_\alpha(x))}(R_\alpha(x)). $$ This shows that the return time for the cross section $C$ is identically equal to $c$. Now it is easy to construct a diffeomorphism $F$ conjugating $\psi$ to $\varphi_\alpha$ up to time change $t\rightarrow ct$. ($F$ maps $C$ to $S^1$.) \medskip Let us turn to the proof of Claim. We shall show it for complex valued functions. A square integrable function $f\in L^2(S^1)$ can be expressed as $$ f=\sum_{k\in{\mathbb Z}}\hat f_k e^{2\pi i kt}, \ \ \ \hat f_k\in{\mathbb C}, \ \ \sum\vert \hat f_k\vert^2<\infty. $$ Notice that $f\in C^\infty(S^1)$ if and only if for any $r>0$, there exists $C_r>0$ such that $\vert \hat f_k\vert\vert k\vert^r<C_r$ for any $k\in Z$. Now the equation $f=g\circ R_\alpha - g+c$ can be read off as $$ \hat g_k=\hat f_k/(e^{2\pi i k \alpha}-1), \ \ \ c=\hat f_0, $$ where $\hat g_k$ is the Fourier coefficients of $g$. Since $\alpha$ is badly approximable and since $$ \vert e^{2\pi ik\alpha}-1\vert\geq C_1\Vert k\alpha\Vert_{S^1} $$ for some $C_1>0$, we have $$ \vert \hat g_k\vert \vert k\vert^r\leq C_1^{-1} \frac{\vert \hat f_k\vert\vert k \vert^r}{\Vert k\alpha\Vert_{S^1}}\leq C_2 \vert \hat f_k\vert \vert k\vert ^\rho\vert k \vert^r\leq C_3. $$ This shows the smoothness of $g$, and the proof of Claim is complete. We shall omit the proof of the other implication of Proposition \ref{p1}. \bigskip Badly approximability of the slope can be defined for Kronecker flows on higher dimensional torus, and these flows are also shown to be parameter rigid. They constitute all the known examples. For related topics, see \cite{AS,dS, LS}. Here is a criterion of the parameter rigidity of a flow, which will be useful for the study of dynamical properties of such flows. \begin{Proposition} \label{p2} The nonsingular flow $\varphi$ on $M$ given by a vector field $X$ is parameter rigid if and only if for any $f\in C^\infty(M)$, there exist $g\in C^\infty(M)$ and $c\in{\mathbb R}$ such that $f=X(g)+c$. \end{Proposition} Let us show the only if part of the proposition. (The other part is just by an analogous argument, using the integration instead of the differentiation.) Choose $f\in C^\infty(M)$. We are going to seek for $g$ and $c$ as in the proposition. It is no loss of generality to assume that $f$ is positive. Let $\psi$ be the flow defined by the vector field $(1/f)X$. Since $\psi$ and $\varphi$ has the same orbit foliation, there is a function $\tau:{\mathbb R}\times M \rightarrow {\mathbb R}$ such that $$ \varphi^t(x)=\psi^{\tau(t,x)}(x). $$ Take $\frac{d}{dt}\vert_{t=0}$ of both sides, and we get $$ \frac{d}{dt}\vert_{t=0}\tau(t,x)=f(x). $$ Now by the parameter rigidity of $\varphi$, there is a diffeomorphism $F$ and $c\in {\mathbb R}$ such that $\psi^{ct}(F(x))=F(\varphi^t(x))$. The diffeomorphism $F$ preserves the orbits, and can be written as $F(x)=\psi^{-g(x)}(x)$ for some $g\in C^\infty(M)$. We have \begin{gather*} \psi^{ct-g(x)}(x)=\psi^{-g(\varphi^t(x))}(\varphi^t(x)), \ \ \rm{ and \ thus} \\ \psi^{g(\varphi^t(x))-g(x)+ct}(x)=\varphi^t(x). \end{gather*} That is, $$ \tau(t,x)=g(\varphi^t(x))-g(x)+ct. $$ Taking $\frac{d}{dt}\vert_{t=0}$, we obtain $f=X(g)+c$, as is desired. \bigskip An important implication of this observation is: \medskip \begin{Corollary} \label{c1} A parameter rigid flow is uniquely ergodic, and it leaves smooth volume form invariant. \end{Corollary} Notice that a flow $\varphi$ is uniquely ergodic if and only if for any $C^\infty$ map $f$, the orbit average $$ \frac{1}{T}\int_0^Tf(\varphi^t(x)dt $$ converges to a constant function, uniformly on $x\in M$. Thus Propositon \ref{p2} immediately implies the unique ergodicity. For the latter statement, let ${\rm vol}$ be a Riemannian volume form on $M$. Take the Lie derivative: $ L_X({\rm vol})=f\cdot{\rm vol} $. By Proposition \ref{p2}, we have $$ L_X({\rm vol})=(X(g)+c){\rm vol}. $$ The integaration over $M$ shows that $c=0$. Thus we have $$ X(e^{-g}\cdot{\rm vol})=0, $$ {\em i.\ e.} $e^{-g}\cdot {\rm vol}$ is an invariant volume form. \bigskip However even among those flows which satisfy the necessary conditions of the previous corollary, no examples of parameter rigid flows are found except the linear flows on higher dimensional torus with badly approximable slopes. Here are two typical negative results in this direction. Let $F_\lambda: T^2 \rightarrow T^2$ be a diffeomorphism defined by $$ F_\lambda(x,y)=(x+y, y+\lambda), $$ where $\lambda$ is a real number, and let $\varphi_\lambda$ be the suspension flow. For irrational $\lambda$, the flow $\varphi_\lambda$ satisfies the conditions of Corollary \ref{c1}. However we have the following result found in \cite{K}. \begin{Theorem} \label{t1} For any real number $\lambda$, the flow $\varphi_\lambda$ is not parameter rigid. \end{Theorem} The other result is about the horocycle flows. Let us introduce them briefly. Let $M$ be the quotient $\Gamma\setminus SL(2;{\mathbb R})$ by a cocompact lattice $\Gamma$. On $M$ the right action of the one parameter subgroup $$ \{ \left[ \begin{array}{cc} 1 & t \\ 0 & 1\end{array} \right] \mid t\in{\mathbb R} \} $$ defines a flow, called {\em horocycle flow}. Horocycle flows are known to satisfy the conditions of Corollary \ref{c1} (\cite{F}). The following theorem is due to L. Fluminio and G. Forni (\cite{FF}). \begin{Theorem} \label{t2} The horocycle flow on a compact manifold $M$ is not parameter rigid. \end{Theorem} We will discuss these two theorems later in the next section after the leafwise cohomology of a foliation is introduced. \smallskip By the way, here is a very simple geometric proof of Theorem \ref{t2} when the manifold $M$ is not a rational homology sphere. Let us take a basis of the Lie algebra $\mathfrak{sl}(2,{\mathbb R})$ as follows. $$ Y=\frac{1}{2} \left[ \begin{array}{cc} 1 & \\ & -1 \end{array} \right] , \ \ S= \left[ \begin{array}{cc} & 1 \\ & \end{array} \right] , \ \ U= \left[ \begin{array}{cc} & \\ 1 & \end{array} \right] $$ These are left invariant vector fields on $SL(2;{\mathbb R})$ and induces vector fields on $M=\Gamma\setminus SL(2;{\mathbb R})$. The horocycle flow $\varphi$ is the one defined by the vector field $S$. Let us denote by $\eta$, $\sigma$ and $u$ the left invariant 1-forms on $SL(2;{\mathbb R})$ which are dual to $Y$, $S$ and $U$. They also induce 1-forms on $M$. Since the first Betti number of $M$ is nonzero, there is a closed 1-form $\omega$ such that the period map $$ [\omega]:\pi_1(M)\rightarrow {\mathbb R} $$ is nontrivial and takes value in ${\mathbb Z}$. Let us write $\omega$ as $$ \omega= f\sigma+*\eta+*u\ \ \ f,*\in C^\infty(M). $$ Assume for contradiction that $f=S(g)+c$. Let $\hat M$ be the cyclic covering of $M$ associated with the homomorphism $[\omega]$. Then the lift $\hat \omega$ of $\omega$ is exact and the primitive $\hat h\in C^\infty(\hat M)$ is proper. Denote by $\hat \varphi$ the lift of the horocycle flow $\varphi$ to $\hat M$. For any $x\in M$, any lift $\hat x$ of $x$, and any $T>0$, we have \begin{gather} \hat h(\hat\varphi^T(\hat x))-\hat h(\hat x) =\int_{\hat\varphi^{[0,T]}(\hat x)}\hat \omega =\int_{\varphi^{[0,T]}(x)}\omega =\int_0^T f(\varphi^t(x))dt \notag \\ =\int_0^T (S(g)(\varphi^t(x))+c)dt =g(\varphi^T(x))-g(x)+cT. \notag \end{gather} Now the lift of the horocycle flow $\hat \varphi$ has a dense orbit (Hedlund). Take $\hat x$ from a dense orbit. Assume $c\neq 0$. Then there exists a sequence $T_i\rightarrow\pm\infty$ such that $\hat \varphi^{T_i}(\hat x)\rightarrow \hat x$. This is a contradiction since the left hand side for $T_i$ tends to 0, while the right hand side tends to $\pm\infty$. Assume now $c=0$. Then since the function $\hat h$ is proper, there exists $T'_j$ such that the left hand side for $T'_j$ tends to $\pm\infty$. Again a contradiction. \smallskip Here is a conjecture by A. Katok (\cite{K}). \begin{Conjecture} An arbitrary parameter rigid flow is smoothly conjugate to a linear flow on the torus of badly approximable slope. \end{Conjecture} This conjecture is known to be true if the manifold has cup length equal to the dimension, or if the manifold is 3-dimensional and has nonvanishing Betti number. \section{Leafwise cohomology and parameter rigidity} In this section, we define the leafwise cohomology of foliations, and discuss its relationship with the parameter rigidity when the foliation is given by a locally free ${\mathbb R}^p$-action. Let $\mathcal F$ be a foliation on a manifold $N$, and denote by $T\mathcal F$ the tangent bundle of $\mathcal F$. A {\em leafwise $k$-form} $\omega$ is a $C^\infty$ cross section of the homomorphism bundle ${\rm Hom}(\bigwedge ^k T{\mathcal F};{\mathbb R})$. The space of leafwise $k$-forms is denoted by $\Omega^k({\mathcal F})$. Thus given $k$ vector fields $X_1$, $X_2$, $\cdots$, $X_k$ tangent to $\mathcal F$ {\em i.\ e.} smooth cross sections of $T\mathcal F$, and $\omega\in \Omega^k(\mathcal F)$, a smooth function $\omega(X_1, X_2, \cdots. X_k)$ is defined. The integrability condition for $T\mathcal F$ allows us to define the {\em leafwise exterior derivative} $$ d_\mathcal F: \Omega^k(\mathcal F)\rightarrow \Omega^{k+1}(\mathcal F) $$ just as the usual exterior derivative. For example, for $\omega\in \Omega^1(\mathcal F)$ we define $$ (d_\mathcal F \omega) (X_1,X_2)=X_1(\omega(X_2))-X_2(\omega(X_1))-\omega([X_1,X_2]). $$ Then $(\Omega(\mathcal F),d_\mathcal F)$ constitutes a cochain complex, whose cohomology is called the {\em leafwise cohomology} of $\mathcal F$, denoted by $H^*(\mathcal F)$. This cochain complex is not elliptic, and the leafwise cohomology can be infinite dimensional. \begin{Example} {\em The 0-dimensional leafwise cohomology $H^0(\mathcal F)$ coincides with the vector space formed by {\em basic functions} {\em i.\ e.} smooth functions on $M$, constant along the leaves. Thus if $\mathcal F$ admits a dense leaf, then $H^0(\mathcal F)\cong{\mathbb R}$.} \end{Example} \begin{Example} \label{ex1} {\em Let $N=L\times T$ and let $\mathcal F$ be the foliation on $N$ whose leaves are $L\times\{ t\}$, $t\in T$. Then we have} $$ H^k(\mathcal F)= H^k(L;C^\infty (T))=H^k(L;{\mathbb R})\otimes C^\infty(T). $$ \end{Example} The cochain space $\Omega^k(\mathcal F)$ is equipped with the Whitney $C^\infty$ topology, and the leafwise exterior derivative $d_\mathcal F$ is continuous. Thus the cocycle space ${\rm Ker}(d_\mathcal F)$ is closed. But the coboundary space ${\rm Im}(d_\mathcal F)$ is not necessarily closed. The quotient of the cocycle space by the closure of the coboundary space is called the {\em reduced leafwise cohomology} and is denoted by $\mathcal H^*(\mathcal F)$. J. \'Alvarez L\'opez and G. Hector (\cite{AH}) have given sufficient conditions for the foliations to have infinite dimensional reduced cohomology, and raised a lot of examples. Their examples are for the most part foliations with dense leaves, definitely not like Example \ref{ex1}. But in this talk, we are mainly interested in such a foliation for which the leafwise cohomology is finite dimensional. \bigskip There are two ways, important for us, to produce elements of the leafwise cohomology. First notice that the restriction map $$ r: {\rm Hom}(\bigwedge^*TM;{\mathbb R})\rightarrow{\rm Hom}(\bigwedge^*T\mathcal F;{\mathbb R}) $$ induces a cochain homomorphism (denoted by the same letter) $$ r:\Omega^*(M)\rightarrow \Omega^*(\mathcal F), $$ where $\Omega^*(M)$ denotes the de Rham complex of $M$. This induces a homomorphism of the cohomology groups $$ r_*: H^*(M;{\mathbb R}) \rightarrow H^*(\mathcal F). $$ The homomorphism $r_*$ is often nontrivial and yields elements of $H^*(\mathcal F)$. Secondly, suppose that the foliation $\mathcal F$ is given by a locally free right action of a Lie group $G$. Then by the differentiation, we get a Lie algebra homomorphism $\iota: \mathfrak g \rightarrow \mathcal X^\infty(\mathcal F)$, where $\mathfrak g=T_eG$ is identified with the Lie algebra of the left invariant vector fields on $G$, and $\mathcal X^\infty(\mathcal F)$ the Lie algebra of the vector fields of $M$ tangent to the foliation $\mathcal F$. Let $X_1,\cdots X_p$ be the basis of $\mathfrak g$. Then a leafwise $k$-form $\omega$ is completely determined by $\omega(\iota X_{i_1},\cdots, \iota X_{i_k})$ for $1\leq i_1 < \cdots < i_k\leq p$, since for each point $x\in M$ the tangent space $T_x(\mathcal F)$ is spanned by $(\iota X_1)_x, \cdots (\iota X_p)_x$. Thus given a left invariant $k$-form $$\omega:\mathfrak g \times\cdots\times \mathfrak g \rightarrow {\mathbb R},$$ a leafwise $k$-form $\iota \omega$ is defined by $$ \iota\omega(\iota X_{i_1},\cdots,\iota X_{i_k}) = \omega(X_{i_1},\cdots, X_{i_k}). $$ This induces a homomorphism $$\iota_*: H^*(\mathfrak g)\rightarrow H^*(\mathcal F).$$ \begin{Proposition}\label{p3} (1) The homomorphism $\iota_*: H^1(\mathfrak g)\rightarrow H^1(\mathcal F)$ is injective. (2) If $G={\mathbb R}^p$, then $\iota_*: H^i({\mathbb R}^p)\rightarrow H^i(\mathcal F)$ is injective for any $i\geq 0$. \end{Proposition} \begin{Remark} The cohomology of the abelian Lie algebra ${\mathbb R}^p$ is isomorphic to the exterior algebra of ${\mathbb R}^p$, and hence to the cohomology of the $p$-torus: $$ H^*({\mathbb R}^p)\cong \bigwedge^*{\mathbb R}^p\cong H^*(T^p;{\mathbb R}). $$ \end{Remark} \noindent Let us give a proof of Proposition \ref{p3}. Let $\xi_1,\cdots, \xi_r$ be the closed left invariant 1-forms whose classes form a basis of $H^1(\mathfrak g)$. They are of course linearly independent in the dual space $\mathfrak g^*$ of $\mathfrak g$. Therefore there exist elements $X_1,\cdots, X_r$ of $\mathfrak g$ such that $\xi_i(X_j)=\delta_{ij}$. Assume $$ a_1\iota\xi_1+\cdots+a_r\iota\xi_r=d_\mathcal Ff $$ for $a_i\in{\mathbb R}$ and $f\in C^\infty(M)$. Let $\gamma_T:[0,T]\rightarrow M$ be an integral curve of the vector field $\sum_ia_i\iota X_i$. Then we have $$ f(\gamma_T(T))-f(\gamma_T(0))=\int_{\gamma_T}\sum_i a_i\iota\xi_i =T\sum_i a_i^2. $$ Since $T$ can be arbitrarily large and the left hand side is bounded, all the coefficients $a_i$'s must vanish. This shows that the classes of $\iota\xi_i$ are linearly independent in $H^1(\mathcal F)$. The proof for the second part is basically similar and use the Stokes theorem on an embedding of the rectangle $[0,T]\times\cdots\times [0,T]$. \bigskip By Proposition \ref{p3}, if $\mathcal F$ is the orbit foliation of a locally free ${\mathbb R}^p$ action, then we have $\dim H^1(\mathcal F)\geq p$. The following proposition relates the leafwise cohomology of the orbit foliation to the parameter rigidity of the action. \begin{Proposition} \label{p4} A locally free effective ${\mathbb R}^p$-action is parameter rigid if and only if its orbit foliation $\mathcal F$ satisfies $\dim H^1(\mathcal F)= p$. \end{Proposition} An action is said to be effective if the isotropy subgroup of some point of the manifold is trivial. The proof for $p=1$ is in order. Let $X$ be the vector field defining a nonsingular flow $\varphi$. Define a leafwise 1-form $\omega$ by $\omega(X)=1$. An arbitrary leafwise 1-form (which is always closed by the dimension reason) is written as $f\omega$ for some $f\in C^\infty(M)$. Given $g\in \Omega^0(\mathcal F)=C^\infty(M)$, notice that $(d_\mathcal F g)(X)=X(g)$, and thus $d_\mathcal F g=X(g)\omega$. Therefore $\dim H^1(\mathcal F)=1$ if and only if for any $f\omega$, there exist $g\in C^\infty(M)$ and $c\in{\mathbb R}$ such that $$ f\omega=X(g)\omega +c\omega, $$ which is equivalent to the parameter rigidity of the flow $\varphi$ by Proposition \ref{p2}. The proof for $p\geq 2$ is found in \cite{MM}. \smallskip Now let us return to the flows in Theorems \ref{t1} and \ref{t2}. In fact what is proven respectively by A. Katok, and L. Flaminio and G. Forni are much stronger than stated there. \begin{Theorem} Let $\mathcal F_\lambda$ be the orbit foliation of the flow $\varphi_\lambda$ in Theorem \ref{t1}. Then the leafwise cohomology $H^1(\mathcal F_\lambda)$ is infinite dimensional. Moreover if $\lambda$ is badly approximable, then we have ${\mathcal H}^1(\mathcal F_\lambda)=H^1(\mathcal F_\lambda)$, {\em i.\ e.} the space of the coboundaries is closed. \end{Theorem} \smallskip \begin{Theorem} Let $\mathcal F$ be the orbit foliation of the horocycle flow. Then $H^1(\mathcal F)$ is infinite dimensional and we have $H^1(\mathcal F)=\mathcal H^1(\mathcal F)$. \end{Theorem} \section{How to compute leafwise cohomology} Here we will show some methods to compute the leafwise cohomology of a foliation. Let $\pi: M \rightarrow B$ be a {\em foliated bundle} {\em i.\ e.} a fiber bundle equipped with a foliation $\mathcal F$ on $M$ which is transverse to each fiber. Thus the dimension of the leaves of $\mathcal F$ is greater than or equal to the dimension of the base $B$. An important class of foliated bundes is obtainded by a construction called suspension. Let $(F,\mathcal G)$ be a foliated manifold. Denote by ${\rm Diff}(F, \mathcal G)$ the group of diffeomorphisms of $F$ which leaves the foliation $\mathcal G$ invariant. Let $h:\Gamma\rightarrow {\rm Diff}(F, \mathcal G)$ be a homomorphism from a group $\Gamma$. Let $B$ be a manifold such that $\pi_1(B) \cong \Gamma$. Then a foliated manifold $(M,\mathcal F)$ called the {\em suspension} of $h$ is defined as follows. On the product $F\times\widetilde B$, where $\widetilde B$ stands for the universal covering of $B$, there is defined a foliation $\widetilde{\mathcal F}$ whose leaves are $L\times \widetilde B$, where $L$ is a leaf of $\mathcal G$. Consider the diagonal action of $\Gamma$ on $F \times\widetilde B$, on the first factor through $h$ and on the second by deck transformation. This action clearly preserves the foliation $\widetilde{\mathcal F}$ and as the quotient we have a foliated manifold $(M, \mathcal F)$. The following fundamental result is due to A. El Kacimi and A. Tihami (\cite{ET}). \begin{Theorem} \label{t3} For the suspension foliation, there is a spectral sequence such that $$ E^{p,q}_2=H^p(B; H^q(\mathcal G)) $$ which converges to $H^{p+q}(\mathcal F)$, where $H^q(\mathcal G)$ is a system of local coefficents on which $\Gamma=\pi_1(B)$ acts through the homomorphism $h$. \end{Theorem} This shows for example that if $H^q(\mathcal G)$ is finite dimensional for any $0\leq q \leq n$, then the leafwise cohomology $H^n(\mathcal F)$ is finite dimensional. \smallskip Let us consider a special case where the foliation $\mathcal G$ on the fiber is a point foliation. In this case we call $\mathcal F$ the {\em suspension of a point foliation} by a homomorphism $h:\Gamma\rightarrow {\rm Diff}(F)$. We have $H^k(\mathcal G)=0$ unless $k=0$. Thus the spectral sequence of Theorem \ref{t3} collapses and we have: \begin {Corollary} \label{c2} If $\mathcal F$ is the suspension of a point foliation by a homomorphism $h:\Gamma=\pi_1(B)\rightarrow {\rm Diff}(F)$ and if $\widetilde B$ is contractible, then we have: $$ H^p(\mathcal F)=H^p(\Gamma; C^\infty(F)), $$ for any $p\geq 0$, where $\Gamma$ acts on $C^\infty(F)$ through the homomorphism $h: \Gamma\rightarrow {\rm Diff}(F)$. \end{Corollary} \bigskip Here is another way for computing the leafwise cohomology, completely different from the above, due to A. Haefliger (\cite{H}). \begin{Theorem} \label{t5} Let $\mathcal F$ be the suspension of a point foliation by a homomorphism $h:\Gamma\rightarrow {\rm Diff}(F)$. Then there is an isomorphism $$ \int_{\mathcal F}:H^{{\rm dim}\mathcal F}(\mathcal F) \rightarrow H_0(\Gamma, C^\infty(F)). $$ \end{Theorem} The group $H_0(\Gamma, C^\infty(F))$ is by definition the quotient space of $C^\infty(F)$ by the subspace spanned by the elements $f\circ h(\gamma) -f$ for $f\in C^\infty(F)$ and $\gamma\in \Gamma$. In sections below, we shall show computational results for some concrete foliations. \section{Linear foliations} Let us consider the suspension foliation of a point foliation, where the group $\Gamma$ is ${\mathbb Z}^p$, the fiber $F$ is $T^q$, and the homomorphism $h$ is given by $$ h:{\mathbb Z}^p\rightarrow T^q \subset {\rm Diff}(T^q), $$ where $T^q$ is to be the group of the translations of $T^q$. Let $B$ be a real valued $p\times q$ matrix. \begin{Definition} {\em The matrix $B$ is called {\em badly approximable} if there exist $C>0$ and $\rho>0$ such that for any $k\in {\mathbb Z}^q\setminus\{0\}$, we have} $$ \Vert Bk \Vert_{T^p}\geq C\vert k \vert^{-\rho}. $$ \end{Definition} Given a matrix $B$, a homomorphism $$h_B:{\mathbb Z}^p\rightarrow {\rm Diff}(T^q)$$ is defined by $$ h_B(n)(x)=x+nB. $$ As the suspension of $h_B$ over the base manifold $T^p$, we get a linear foliation $\mathcal F_B$ on $T^{p+q}$. The matrix $B$ is called the {\em slope matrix} of the foliation $\mathcal F_B$. \begin{Theorem}{\bf \cite{AS}} \label{t4} If the slope matrix $B$ is badly approximable, then the homomorphism $$ \iota_*: H^i({\mathbb R}^p)\rightarrow H^i(\mathcal F_B) $$ is an isomorphism for any $i\geq 0$. \end{Theorem} Recalling the definition of the homomorphism $\iota_*$, we get the following corollary, which will be useful in Sect.\ 11. \begin{Corollary} \label{c3} Under the same assumption as above, the homomorphism $$ r_*: H^i(T^{p+q})\rightarrow H^i(\mathcal F_B) $$ is a surjection for any $i\geq 0$. \end{Corollary} As is easily shown the foliation $\mathcal F_B$ is the orbit foliation of an ${\mathbb R}^p$ action. Since $H^1(\mathcal F_B)\cong {\mathbb R}^p$, this action is parameter rigid if the action is effective. In fact a bit wider class of ${\mathbb R}^p$-actions for which $\iota_*$ is an isomorphism are found in \cite{Lu}. \section{Orbit foliation of transversely hyperbolic ${\mathbb R}^p$ actions} Abundant examples of $C^1$ locally rigid (hence parameter rigid) ${\mathbb R}^p$-actions are presented by A. Katok and others (\cite{K, HK, KL,KS1,KS2}). Here let us mention only one type among them. Let us denote by $Aff_+(T^{p+1})$ the group of the orientation preserving affine transformations of $T^{p+1}$, with respect to the standard affine structure of $T^{p+1}$. Let us consider a homomorphism $$ h: {\mathbb Z}^p\rightarrow Aff_+(T^{p+1}) $$ with the following property; \smallskip \noindent (*) {\em The derivative $h_*:{\mathbb Z}^p\rightarrow SL(p+1;{\mathbb Z})$ is injective and the image is generated by hyperbolic elements.} \smallskip In this case, the image $h_*({\mathbb Z}^p)$ is shown to be simultaneously diagonalizable in $SL(p+1;{\mathbb R})$. \begin{Theorem} {\bf \cite{KL}} If $p\geq 2$, the cohomology $H^1({\mathbb Z}^p; C^\infty(T^{p+1}))$ associated to $h$ above is isomorphic to ${\mathbb Z}^p$. \end{Theorem} The suspension foliation $\mathcal F_h$ of $h$ over $T^p$ is the orbit foliation of an ${\mathbb R}^p$ action $\varphi_h$. The above theorem implies that $H^1(\mathcal F_h) \cong{\mathbb R}^p$, via Corollary \ref{c2}. That is, by Proposition \ref{p4}, this action is parameter rigid. Moreover it is shown to be $C^1$ locally rigid (\cite{KL}). \medskip As for the higher cohomology group, however, this foliation does not exhibit such rigid property as the foliation $\mathcal F_B$ does in Theorem \ref{t4}. \begin{Proposition} \label{p5} For $p\geq 1$, the top dimensional cohomology $H^p(\mathcal F_h)$ of the foliation $\mathcal F_h$ is infinite dimensional and we have $\mathcal H^p(\mathcal F_h) =H^p(\mathcal F_h)$. \end{Proposition} The proof uses Theorem \ref{t5}. For $p=1$ this result is classical. For the related topics for Anosov flows, see \cite{L,GK,dLMM}. \bigskip Let us state a rigidity result about hyperbolic ${\mathbb R}^p$ actions, which is global in nature. An ${\mathbb R}^p$ action $\varphi$ on a closed $(2p+1)$-dimensional manifold $M$ is called {\em split hyperbolic} if there is a continuous splitting of the tangent bundle $$ TM=T{\mathbb R}^p\oplus E_1\oplus\cdots\oplus E_{p+1}, $$ where $T{\mathbb R}^p$ is the tangent bundle of the orbit foliation and each $E_i$ is a 1-dimensional subbundle invariant under the differential of the action. Furtheremore we assume that there exist $\xi_1,\cdots, \xi_{p+1}\in{\mathbb R}^p$ such that the flow $\varphi(\cdot,t\xi_i)$ is expanding along $E_i$ and contracting along $E_j$ ($j\neq i$). \begin{Theorem} {\bf \cite{M}} A split hyperbolic ${\mathbb R}^p$ action is $C^\infty$ conjugate to the suspension action $\varphi_h$ of some affine representation $$ h:{\mathbb Z}^p\rightarrow Aff_+(T^{p+1}) $$ with the property (*), up to an automorphism of ${\mathbb R}^p$. \end{Theorem} \section{K\"unneth formula} We shall interrupt the computation of the leafwise cohomology and state a general theorem, which has an interesting application for the parameter rigidity of ${\mathbb R}^n$ actions. A. El Kacimi and A. Tihami (\cite{ET}) followed the arguments of R. Bott and L. W. Tu (\cite{BT}) in the framework of leafwise forms, to obtain {\em e.\ g.} the Mayer-Vietoris theorem or the spectral sequence theorem (Theorem \ref{t3}). One can further pursue this line to obtain the following K\"unneth formula for the leafwise cohomology. Let $\mathcal F$ (resp.\ $\mathcal G$) be a foliation on a manifold $M$ (resp.\ $N$). Then on the product manifold $M\times N$, products of leaves of $\mathcal F$ and $\mathcal G$ form a foliation called the {\em product foliation}, denoted by $\mathcal F\times \mathcal G$. \begin{Theorem} \label{12} Assume $\dim H^j(\mathcal G)<\infty$ for $0\leq j \leq k$. Then we have an isomorphism: $$ \sum_{i+j=k}H^i(\mathcal F)\otimes H^j(\mathcal G)\cong H^k(\mathcal F\times \mathcal G). $$ \end{Theorem} Here is an immediate corollary concerning ${\mathbb R}^n$ actions.. Given two actions $$ \varphi:M\times{\mathbb R}^p\rightarrow M \ \ {\rm and} \ \ \psi:N\times{\mathbb R}^q\rightarrow N, $$ an ${\mathbb R}^{p+q}$ action $\varphi\times\psi$ on $M\times N$ called the {\em product action} is defined by $$ (\varphi\times\psi)((x,y),(s,t))=(\varphi(x,s),\psi(y,t)), $$ where $x\in M$, $y\in N$, $s\in {\mathbb R}^p$ and $t\in{\mathbb R}^q$. \begin{Theorem} Let $M$ and $N$ be closed manifolds. Assume that $\varphi$ and $\psi$ are parameter rigid actions of connected abelian Lie groups and that the product action $\varphi\times\psi$ is effective. Then $\varphi\times\psi$ is parameter rigid. \end{Theorem} For example the products of the linear actions of Sect.\ 5 and the transversely hyperbolic actions of Sect.\ 6 can be parameter rigid. It seems that this theorem is difficult to prove without resorting to an algebraic topological argument. Now in the next two sections, we will be back to the computation of the leafwise cohomology. \section{The weak stable foliation of the suspension Anosov flow} Let $\mathcal W^s$ be the weak stable foliation of the suspension flow of a hyperbolic toral automorphism $A: T^n\rightarrow T^n$. Let $\mathcal V^s$ be the linear foliation on $T^n$ whose leaves are parallel translations of the stable eigenspace $E^s$. The automorphism $A$ leaves $\mathcal V^s$ invariant, and as is well known the slope matrix of the linear foliation $\mathcal V^s$ is badly approximable. By Theorem \ref{t4} we have $$ H^*(\mathcal V^s)\cong H^*({\mathbb R}^p)\cong\bigwedge^*{\mathbb R}^p, $$ where $p$ denotes the dimension of the foliation $\mathcal V^s$. The foliation $\mathcal W^s$ in question is the suspension of $(T^n, \mathcal V^s)$ by a homomorphism $h:{\mathbb Z}\rightarrow {\rm Diff}(T^n,\mathcal V^s)$ given by $h(1)=A$. In the case of foliated bundles over $S^1$, Theorem \ref{t3} gives rise to the Wang exact sequence; $$ H^{k-1}(\mathcal V^s) \rightarrow H^{k-1}(\mathcal V^s) \rightarrow H^k(\mathcal W^s) \rightarrow H^k(\mathcal V^s) \rightarrow H^k(\mathcal V^s). $$ The first and the last arrows are $A^*-{\rm Id}$. Notice that $A^*:H^i(\mathcal V^s)\rightarrow H^i(\mathcal V^s)$ is the $i$-th exterior product of $A\vert_{E^s}$, under the identification given by $\iota_*$: $\bigwedge^i E^s\cong H^i({\mathbb R}^p) \cong H^i(\mathcal V^s)$. Thus we get: \begin{Proposition} \label{pa1} $$ H^i(\mathcal W^s)={\mathbb R} \ {\rm for} \ i=0,1, \ \ {\rm and} \ \ H^i(\mathcal W^s)=0 \ {\rm for} \ i\geq 2. $$ \end{Proposition} We shall return to this result in Sect.\ 11. \section{Stable foliations of geodesic flows} There are foliations for which neither Corollary \ref{c2} nor Theorem \ref{t5} apply but still we can compute the leafwise cohomology by a geometric method. Let $M=\Gamma\setminus {\rm SL}(2;{\mathbb R})$ where $\Gamma$ is a cocompact lattice and let $\mathcal F^s$ be the foliation defined by the action of the subgroup $$ \{ \left[ \begin{array}{cc} e^{t/2} & x \\ 0 & e^{-t/2}\end{array} \right] \mid t,x\in{\mathbb R} \}. $$ This subgroup is isormorphic to the Lie group $GA$ of the orientation preserving affine action of the real line and we have $H^1(\mathfrak g\mathfrak a)\cong{\mathbb R}$. The foliation $\mathcal F^s$ is the weak stable foliation of an Anosov flow. \begin{Theorem} \label{t6} The homomorphism $$ (r_*,\iota_*): H^1(M;{\mathbb R})\oplus{\mathbb R}\rightarrow H^1(\mathcal F^s) $$ is an isomorphism. \end{Theorem} Since there are many \lq\lq leafwise 1-currents" given by closed orbits of the Anosov flow, it is not difficult to show that $(r_*,\iota_*)$ is injective. To show the surjectivity, we use the following lemma. \begin{Lemma} Given a $C^\infty$ function $k$, the equation \begin{equation} \label{e2} k=Y(g)+g \end{equation} has a $C^\infty$ solution $g$ if and only if there exists a $C^\infty$ function $l$ fulfilling $$ S(k)=Y(l). $$ \end{Lemma} Here $Y$, $S$ and $U$ are the vector fields on $M$ induced by the action of the elements of the Lie algebra ${\mathfrak sl}(2;{\mathbb R})$; $$ Y=\frac{1}{2} \left[ \begin{array}{cc} 1 & 0 \\ 0 & 1\end{array} \right] , \ \ S= \left[ \begin{array}{cc} 0 & 1 \\ 0 & 0\end{array} \right] , \ \ U= \left[ \begin{array}{cc} 0 & 0 \\ 1 & 0 \end{array} \right] . $$ The integration along the orbit of the $Y$-flow from $t=-\infty$ gives a continuous solution $g$ of (\ref{e2}), which is $C^\infty$ along $Y$ and $U$ directions. The condition $S(k)=Y(l)$ transforms the difference of $g$ between nearby two points on an $S$-leaf to the sum of the difference of the images of these points by time $-T$ of the $Y$-flow and the integration of $k$ along this long $S$-orbit. Then the unique ergodicity of $S$-flow (\cite{F}) ensures that $g$ is $C^\infty$ along the $S$-direction. \medskip Here is also a result about 2-dimensional cohomology, which is a translation of a result of A. Haefliger found in \cite{H}. An application of Theorem \ref{t5}. \begin{Theorem} \label{t7} The homomorphism $r_*: H^2(M;{\mathbb R})\rightarrow \mathcal H^2(\mathcal F^s)$ is an isomorphism. \end{Theorem} Unfortunately the above theorem is only about the reduced leafwise cohomology $\mathcal H^2(\mathcal F^s)$. The honest leafwise cohomology $H^2(\mathcal F^s)$ is not known. We will be back to this problem in Sect.\ 11. \section{Parameter rigidity of certain solvable Lie group actions} First of all let us state a criterion for the parameter rigidity of the action of a general Lie group, which takes form of a nonlinear equation if the Lie group is nonabelian. Let $\varphi:M\times G\rightarrow M$ be a locally free action of a connected and simply connected Lie group $G$ with Lie algebra $\mathfrak g$. Denote by $\mathcal F$ the orbit foliation. As is explained in Sect.\ 3, the vector field $\iota(X) \in T\mathcal F$ is defined for each $X\in\mathfrak g$. A $\mathfrak g$-valued leafwise 1-form $\omega_0$ associted with the action $\varphi$ is defined by $$ \omega_0(\iota(X))=X. $$ It satisfies $$ d_{\mathcal F}\omega_0+[\omega_0\wedge\omega_0]=0, $$ where $[\omega_0\wedge\omega_0]$ is a leafwise 2-form defined by $$ [\omega_0\wedge\omega_0](X,Y)=[\omega_0(X),\omega_0(Y)]. $$ \begin{Proposition} \label{p6} Assume the action $\varphi$ is effective. The action $\varphi$ is parameter rigid if for any $\mathfrak g$-valued leafwise 1-form $\omega$ such that $$ d_{\mathcal F}\omega+[\omega\wedge\omega]=0, $$ there exist an endomorphism $\Phi$ of $G$ and a smooth map $b: M\rightarrow G$ such that $$ \omega=b^{-1}\Phi_*(\omega_0)b+b^{-1}db. $$ \end{Proposition} As an application let us consider the foliation of Sect.\ 8. Computation based on the result of Theorem \ref{t6} leads to an alternative proof of the following theorem by E. Ghys (\cite{G}). \begin{Theorem} \label{11} The action of the Lie group $GA$ in Sect.\ 8 is parameter rigid, if the manifold $M$ is a rational homology sphere. \end{Theorem} E. Ghys also showed that if the manifold $M$ has nonvanishing first Betti number, then the acion of $GA$ is parameter rigid among those actions which preserve the volume form. Our method using the leafwise cohomology just yields the same conclusion. Nothing more, except we can show that the sufficient condition of Proposition \ref{p6} are not satisfied in this case. So the question of the parameter rigidity of the $GA$-actions for 3-manifolds with nonvanishing first Betti number is still open. \medskip Let us raise other examples. Let $A$ be a hyperbolic automorphism of the torus $T^n$ ($n\geq 2$) and let $\mathcal W^s$ be the weak stable foliation of the suspension flow. By Proposition \ref{pa1}, we have $H^1(\mathcal W^s)\cong{\mathbb R}$. If the matrix $A$ has no negative eigenvalues of absolute value smaller than 1, then the foliation $\mathcal W^s$ is the orbit foliation of a locally free action of a two step solvable group $G$. \begin{Theorem} {\bf \cite{MM}} If the characteristic polynominal of $A$ is irreducible over ${\mathbb Q}$, then the $G$ action is parameter rigid. \end{Theorem} Together with the rigidity result of the orbit foliations established by A. El Kacimi and M. Nicolau (\cite{EN}), we obtain the following. \begin{Theorem} If furthermore all the eigenvalues of $A$ are positive, then the $G$ action is $C^\infty$ rigid. \end{Theorem} \section{Minimizable foliation} A foliation $\mathcal F$ on a manifold $M$ is called {\em minimizable} if there is a Riemannian metric on $M$ for which all the leaves of $\mathcal F$ are minimal surfaces. \begin{Proposition} A foliation $\mathcal F$ on the total space $E$ of a fiber bundle $\pi:E\rightarrow B$ transverse to each fiber and of complementary dimension is minimizable. \end{Proposition} In fact a Riemannian metric on $B$ can be lifted to a leafwise Riemannian metric of $\mathcal F$ and extended to a Riemannian metric of $E$ in such a way that each fiber is orthogonal to each leaf. Clearly any leaf is totally geodesic for this metric. Thus most of the examples of foliations so far listed are minimizable. However there is an exception. \begin{Proposition} Horocycle flows on closed 3-manifolds are not minimizable. \end{Proposition} In fact a foliation having a positive holonomy invariant measure which is the boundary of an invariant 1-current is not minimizable (\cite{S}). Here is a criterion for the minimizability due to H. Rummler and D. Sullivan (\cite{R,S}). Assume all the foliations in this section are oriented. \begin{Theorem} \label{13} Let $g_0$ be a leafwise Riemannian metric of a $p$-dimensional foliation $\mathcal F$. It extends to a Riemannian metric of the whole manifold $M$ for which all the leaves are minimal surfaces if and only if the leafwise volume form $\omega_0$ induced by $g_0$ extends to a form $\omega$ of $M$ such that $$ d\omega(\xi_1,\cdots,\xi_{p+1})=0 $$ whenever the first $p$ of the vector fields $\xi_i$'s are tangent to the leaves. \end{Theorem} Now we call a foliation $\mathcal F$ {\em totally minimizable} if any leafwise Riemannian metric extends to a Riemannian metric for which all the leaves are minimal. \begin{Corollary} \label{14} A $p$-dimensional foliation $\mathcal F$ on $M$ is totally minimizable if the homomorphism $$ r_*:H^p(M;{\mathbb R})\rightarrow H^p(\mathcal F) $$ is surjective. If ${\rm codim}\mathcal F=1$, then the converse is also true. \end{Corollary} In fact if $r_*$ is surjective, then the leafwise volume form $\omega_0$ of any leafwise Riemannian metric can be expressed as $$ \omega_0=r(\omega)+d_{\mathcal F}(\eta_0), $$ for a closed $p$-form $\omega$ and a leafwise $(p-1)$-form $\eta_0$. But since $r:\Omega^{p-1}(M)\rightarrow\Omega^{p-1}(\mathcal F)$ is surjective, we have $\eta_0=r(\eta)$ for some $(p-1)$-form $\eta$. It follows then that $$ \omega_0=r(\omega+d\eta), $$ for a closed $p$-form $\omega+d\eta$, whence the criterion of Theorem \ref{13} is satisfied. To show the converse statement for codimension one foliation, notice first of all that since $\dim(M)=p+1$, the condition for the $(p+1)$-form $d\omega$ in Theorem \ref{13} is equivalent to saying that $d\omega=0$. Next notice that any leafwise $p$-form is the difference of the leafwise volume forms of two leafwise Riemannian metrics. \medskip As applications, Corollary \ref{c3} implies that the linear foliations on the torus with badly approximable slope matrices are totally minimizable. Also the foliation $\mathcal W^s$ of Sect.\ 8. is totally minimizable by Proposition \ref{pa1}. \medskip A foliation $\mathcal F$ is said to be {\em almost totally minimizable} if any leafwise Riemannian metric is approximated arbitrarily close in the $C^\infty$ topology by the restriction of Riemannian metrics for which all the leaves are minimal. One example of such foliations are $p$-dimensional linear foliations $\mathcal F$ on the torus $T^n$ whose leaves are all dense. This is a consequence of the surgectivity of the map $r_*:H^p(T^n)\rightarrow \mathcal H^p(\mathcal F)$. This result is best possible if the slope matrix of the foliation $\mathcal F$ is not badly approximable. (Otherwise $\mathcal F$ is totally minimizable, as we mentioned above.) Likewise for the foliation $\mathcal F^s$ of Sect.\ 9, the homomorphism $r_*: H^2(M;{\mathbb R})\rightarrow \mathcal H^2(\mathcal F^s)$ is an isomorphism (Theorem \ref{t7}). This implies that the foliation $\mathcal F^s$ is almost totally minimizable (\cite{H,HL}). Unlike the previous example of linear flows, we do not know if this is best possible or not. It is known that the subspaces $\{Y(f)\vert f\in C^\infty(M)\}$ and $\{S(f)\vert f\in C^\infty(M)\}$ are closed in $C^\infty(M)$ (\cite{GK} or Proposition \ref{p5}, and \cite{FF} or Theorem \ref{t2}). The problem is equivalent to deciding if the sum of these two subspaces is closed.
\section{Introduction} Persistent currents in normal-metal rings threaded by an Aharonov-Bohm flux constitute a paradigm of quantum-coherence effects in the thermodynamic properties of mesoscopic systems. While the history of persistent currents dates back to the early days of quantum mechanics \cite{Hund38} and of superconductivity,\cite{superconductivity} they were studied intensively starting with the seminal paper by Buttiker, Imry, and Landauer.\cite{Buttiker83} Most experiments to date detected persistent currents using SQUIDs (superconducting quantum interference device) as magnetometers. A different technique was recently developed by Bleszynski-Jayich {\em et al.}\cite{Harris09} which is much more sensitive and allows for precision measurements with lower back action and over a wider range of magnetic fields. The high-precision cantilever torque magnetometer relies on the interaction of the magnetic moment associated with the persistent current and a large applied magnetic field. This interaction shifts the resonance frequency of a microcantilever on which the rings are located. Measurements of the frequency shift allow one to extract the persistent current quantitatively. This paper extends the existing theory of mesoscopic persistent currents to include a large applied magnetic field. We focus on metallic samples with diffusive electron dynamics for which the applied magnetic fields are {\it non}-quantizing. Our results hold for both normal-metal rings as well as rings made of nominally superconducting materials (provided that the magnetic field significantly exceeds the superconducting critical field $H_{c2}$) and include the effects of spin, namely Zeeman splitting and spin-orbit scattering. Within an independent-electron model, the flux-periodic persistent current is strongly sample specific with both magnitude and sign depending on the details of the disorder configuration and the geometry of the ring. As a result, its ensemble average $\langle I\rangle$ is, even in the canonical ensemble,\cite{Altshuler91,Schmid91,Oppen91} small compared to its second moment $\langle I^2\rangle$ \cite{Cheung89,Riedel93} so that the latter describes the typical magnitude of the persistent current of an individual ring. The typical persistent current is $\phi_0$-periodic and, in a diffusive metallic ring, has an amplitude of the order of $e/\tau_D$, where $\tau_D$ denotes the diffusion time of an electron around the ring. The disorder-averaged persistent current is dominated by the contribution of electron-electron interactions \cite{Ambegaokar90} and is $\phi_0/2$-periodic. In a normal-metal ring, it is of the order of $\lambda(e/\tau_D)$, where $\lambda$ is an effective electron-electron coupling constant. While $\lambda$ is of order unity in lowest order perturbation theory, higher-order contributions are expected to reduce its magnitude significantly. \cite{Altshuler81,Eckern91} In rings made of superconducting materials, a related mechanism leads to a current due to superconducting fluctuations above the critical temperature $T_c$.\cite{Eckern90, Oppen92} These theoretical expectations have been tested in several experiments, including metallic, \cite{Levy90, Chandrasekhar91, Jariwala01, Moler09, Harris09} semiconducting,\cite{Mailly93} as well as superconducting\cite{Moler07} rings. While results of early experiments with metallic rings were in apparent strong disagreement with theoretical predictions, a more recent SQUID-based experiment \cite{Moler09} yielded data reasonably close to theory. Finally, the measurement of the typical persistent current reported in Ref.\ \onlinecite{Harris09} agrees, without any adjustable parameters and over a wide range of experimental variables, with the predictions of the model of noninteracting diffusive electrons described here. We also note in passing that there is a closely related set of works, both experimental and theoretical, which explores the magnetic response of singly-connected mesoscopic systems, see e.g., Refs.\ \onlinecite{Levy93, Oppen94, Ullmo95, Ullmo98}. Persistent currents have also motivated a multitude of further theoretical considerations. Among other results, it was suggested that the persistent current is highly sensitive to a variety of subtle effects, including the coupling of the ring to its electromagnetic environment \cite{Yudson93, Aronov93} as well as magnetic impurities within the ring.\cite{Bary08, Schwiete09} This indicates that accurate measurements and understanding of persistent currents in various settings would address a number of interesting questions in many-body condensed matter physics. This paper is organized as follows. In Sec.\ \ref{periodicity}, we discuss the flux dependence of the persistent current. Sec.\ \ref{independent} discusses the effects of the strong magnetic field on the persistent current within the independent-electron model, including the effects of the Zeeman energy and spin-orbit scattering. Sec.\ \ref{interaction} focuses on the interaction contribution to the persistent current. Sec.\ \ref{datacomparison} contains a detailed comparison between our theoretical results and the experimental data of Ref.\ \onlinecite{Harris09}. We conclude in Sec.\ \ref{conclusions}. \section{Flux periodicity of the persistent current} \label{periodicity} Conventionally, persistent currents are discussed in the limit of a pure Aharonov-Bohm flux threading the ring. In this case, gauge invariance implies flux periodicity,\cite{Byers} \begin{equation} I(\phi)=I(\phi+\phi_0), \end{equation} where the period is given by the flux quantum $\phi_0=h/e$, and time-reversal invariance gives the relation \begin{equation} I(\phi)=-I(-\phi). \end{equation} As a result, the persistent current vanishes at integer and half-integer multiples of the flux quantum and can be expressed as a Fourier series $I(\phi)=\sum_{p=1}^\infty I_p \sin(2\pi p \phi/\phi_0)$. It is instructive to deduce the consequences of this Fourier decomposition for the current-current correlation function \begin{equation} C(\phi,\phi')=\langle I(\phi)I(\phi')\rangle. \end{equation} Here, $\langle\ldots\rangle$ denotes a disorder average. We anticipate that the Fourier components $I_p$ are mutually uncorrelated, i.e., $\langle I_p I_{p'}\rangle = \langle I_p^2\rangle \delta_{pp'}$. Then, we obtain \begin{widetext} \begin{eqnarray} C(\phi,\phi') &=& \sum_{p=1}^\infty \langle I_p^2\rangle \sin(2\pi p \phi/\phi_0) \sin(2\pi p \phi'/\phi_0) = \sum_{p=1}^\infty \frac{\langle I_p^2\rangle}{2} \left\{ \cos[2\pi p (\phi-\phi')/\phi_0] - \cos[2\pi p (\phi+\phi')/\phi_0]\right\}. \end{eqnarray} \end{widetext} Within the diagrammatic approach to diffusive electronic systems, the two terms depending on $(\phi-\phi')$ and $(\phi+\phi')$ have immediate interpretations as the diffuson and cooperon contributions, respectively.\cite{Riedel93} Both contributions are of the same magnitude but depend differently on the magnetic flux. In the presence of an additional large magnetic field $B$ penetrating the metal ring, one expects that the cooperon contribution is strongly suppressed. This leads to a change in the flux dependence of $C(\phi,\phi^\prime)$ which can also be obtained directly from symmetry considerations. While gauge invariance and hence the flux periodicity persist, the additional magnetic field changes the time-reversal relation into $I(B,\phi)=-I(-B,-\phi)$. As a result, the current is no longer odd in the Aharonov-Bohm flux $\phi$ alone, and the Fourier series takes the more general form (at fixed $B$) \begin{equation} I(\phi) = \sum_{p=1}^\infty \{I^{(+)}_p \cos(2\pi p\phi/\phi_0)+I^{(-)}_p \sin(2\pi p\phi/\phi_0)\}. \label{FourierDecomposition} \end{equation} If we again anticipate that the Fourier components are mutually uncorrelated, \begin{eqnarray} \langle I^{(\pm)}_p I^{(\pm)}_{p'}\rangle &=& \langle [I^{(\pm)}_p]^2\rangle \delta_{pp'} \label{Correlators1}\\ \langle I^{(+)}_pI^{(-)}_{p'}\rangle &=& 0 \label{Correlators2} \end{eqnarray} and that, moreover, $\langle [I^{(+)}_p]^2\rangle = \langle [I^{(-)}_p]^2\rangle $, we find \begin{widetext} \begin{eqnarray} C(\phi,\phi') &=& \sum_{p=1}^\infty \langle [I^{(+)}_p]^2\rangle [ \sin(2\pi p \phi/\phi_0) \sin(2\pi p \phi'/\phi_0) + \cos(2\pi p \phi/\phi_0) \cos(2\pi p \phi'/\phi_0) ] \nonumber \\ & =& \sum_{p=1}^\infty \langle [I^{(+)}_p]^2\rangle \cos[2\pi p (\phi-\phi')/\phi_0] . \label{HarmonicContent} \end{eqnarray} \end{widetext} In agreement with expectations, our analysis implies that in the presence of a large magnetic field $B$, the current-current correlation function has the flux dependence of a diffuson contribution. Note that the magnitude of the persistent current, $\langle I^2(\phi) \rangle$, becomes independent of flux. As a special case, this also implies that the persistent current can be nonzero at zero flux. It is interesting to note that in the presence of a large magnetic field, the flux dependence of the persistent current can also be written as \begin{equation} I(\phi) = \sum_{p=1}^\infty I_p \cos(2\pi p \phi/\phi_0 - \alpha). \label{RandomPhase} \end{equation} Comparing with Eq.\ (\ref{FourierDecomposition}) yields the identities $I_p^{(+)} = I_p \cos\alpha$ and $I_p^{(-)} = I_p \sin\alpha$. Then, we automatically reproduce Eqs.\ (\ref{Correlators1}) and (\ref{Correlators2}) by assuming that the phase offset $\alpha$ has a uniform distribution over the disorder ensemble. This also yields the relation $\langle I_p^2\rangle = 2\langle [I^{(\pm)}_p]^2\rangle $. In the next section, we verify these flux dependencies explicitly within the model of diffusive non-interacting electrons. \section{Independent-electron contribution} \label{independent} \subsection{Current-current correlation function} \begin{figure}[t] \includegraphics[width=0.8\columnwidth]{correlator.pdf} \caption{(a) Diffuson and (b) cooperon diagrams for the autocorrelation function of the density of states. Full lines represent electronic Green functions, dashed lines correspond to disorder scattering. \label{diagrams}} \end{figure} The persistent current is obtained as the flux-derivative of the thermodynamic potential \begin{equation} I=-\frac{\partial\Omega}{\partial \phi}. \end{equation} For non-interacting electrons, the (grand-canonical) thermodynamic potential $\Omega$ can be expressed as \begin{equation} \Omega(\mu,{\bf B}) = -T \int dE\, \nu(E,{\bf B}) \ln[1+e^{-\beta(E-\mu)}] \end{equation} in terms of the density of states $\nu(E,{\bf B})$. Here, $\beta=1/T$ denotes the inverse temperature. (We use units $k_B=1$ and $\hbar=1$.) Here, the magnetic field ${\bf B}$ includes both the Aharonov-Bohm flux $\phi$ threading the ring and the magnetic field penetrating the ring. For definiteness, we will from now on decompose the full magnetic field into a pure Aharonov-Bohm contribution and an in-plane field $B_\parallel$ penetrating the ring. Accordingly, we will drop the vector nature of ${\bf B}$ in the following although it should be kept in mind that in principle, the persistent current is not an isotropic function of magnetic field. The thermodynamic potential $\Omega(\mu,B)$ at finite temperature can be related to its zero-temperature limit \begin{equation} \Omega_0(\mu,B) = \int_{-\infty}^\mu dE (E-\mu) \nu(E,B) \end{equation} as \begin{equation} \Omega(\mu,B) = \int_{-\infty}^\infty dE \left(-\frac{\partial f_\mu(E)}{\partial E}\right) \Omega_0(E,B), \end{equation} in terms of the Fermi-Dirac distribution $f_\mu(E)$. Thus, the current-current correlation function $C_I(B,B')=\langle I(B)I(B')\rangle$ takes the form \begin{widetext} \begin{eqnarray} C_I(B,B') &=& \int dE\, dE' \left(-\frac{\partial f_\mu(E)}{\partial E}\right) \left(-\frac{\partial f_\mu(E')}{\partial E'}\right) \frac{\partial^2}{\partial \phi\partial \phi'} \,\langle \Omega_0(E,B) \Omega_0(E',B') \rangle \nonumber\\ &=& \int d\epsilon \,\partial_\epsilon^2 \left(\frac{\epsilon}{1-\exp(-\beta\epsilon)}\right) \frac{\partial^2}{\partial \phi\partial \phi'} \,\langle\Omega_0(E,B)\Omega_0(E',B')\rangle \nonumber\\ &=& \int d\epsilon \,\partial_\epsilon^2 \left(\frac{\epsilon}{1-\exp(-\beta\epsilon)}\right)\,C_I^{(0)} (E,B;E',B') \label{FiniteTemperature} \end{eqnarray} Here, we used in the second identity that the correlator depends only on the energy difference $\epsilon=E-E'$ so that we can perform the integral over the sum $\sigma=E+E'$. Thus, we are led to consider the zero-temperature autocorrelation function $C_I^{(0)} (E,B;E',B') = \langle I(E,B) I(E',B')\rangle$ of currents at different chemical potentials $E$ and $E'$ as well as fields $B$ and $B'$. Within a model of non-interacting, diffusive electrons, the calculation of \begin{equation} C_I^{(0)} (E,B;E',B') = \int_{-\infty}^E dE_1 \int_{-\infty}^{E'} dE_2 (E_1-E)(E_2-E')\frac{\partial^2}{\partial \phi\partial \phi'} \langle \nu(E_1,B)\nu(E_2,B')\rangle \end{equation} starts from the familiar diagrams in Fig.\ \ref{diagrams} for the disorder-averaged autocorrelation function of the density of states.\cite{Shklovskii} Note that {\it both} the diffuson and the cooperon diagram contribute to the persistent current. The diffuson diagram depends on the difference ${\bf A}_-={\bf A}-{\bf A}'$ of the magnetic vector potentials, the cooperon diagram on the sum ${\bf A}_+={\bf A}+{\bf A}'$. Performing the integrations over the fast Green-function arguments in the diagrams of Fig.\ \ref{diagrams}, one arrives at the expression \begin{eqnarray} C_I^{(0)}(E,B;E',B') = \frac{1}{2\pi^2}\frac{\partial^2}{\partial\phi\partial \phi'} {\rm Re} \sum_{\pm}\int_0^\infty d\sigma \int_{-2\sigma}^{2\sigma} d\epsilon \left[\sigma^2 - \frac{\epsilon^2}{4} \right] {\rm Tr}\left(\frac{1}{-D[\nabla-{ie}{\bf A}_\pm]^2 + i(\epsilon+E-E')}\right)^2. \end{eqnarray} Rewriting the square of the diffusion pole as a derivative with respect to $\epsilon$ and integrating by parts yields \begin{eqnarray} C_I^{(0)}(E,B;E',B') = -\frac{1}{4\pi^2}\frac{\partial^2}{\partial\phi\partial \phi'} \sum_{\pm}\int_0^\infty d\sigma {\rm Im} \int_{-2\sigma}^{2\sigma} d\epsilon\, \epsilon {\rm Tr}\left(\frac{1}{-D[\nabla-{ie}{\bf A}_\pm]^2 + i(\epsilon+E-E')}\right). \label{CurrentThroughDifusion} \end{eqnarray} \end{widetext} Here, $D$ denotes the diffusion constant and we limit attention to spinless systems. (Effects of spin will be discussed separately in Sec.\ \ref{spin}.) In Eq.\ (\ref{CurrentThroughDifusion}), the trace is over a space of wavefunctions $\psi$ satisfying the condition \begin{equation} \left. {\bf \hat n}\cdot[\nabla-{ie}{\bf A}_\pm]\psi\right|_\Sigma =0. \label{BoundaryCondition} \end{equation} at the surface $\Sigma$ of the metallic ring. (${\bf \hat n}$ denotes denotes a unit vector normal to the surface.) In general, this boundary condition makes the evaluation of Eq.\ (\ref{CurrentThroughDifusion}) a tedious problem. To simplify this problem, we use a model in which the in-plane magnetic field is taken to be of constant magnitude and to point along the azimuthal direction around the ring. While this toroidal-field model is clearly different from experimental realizations, we expect that it gives a qualitatively and, for certain quantities, even quantitatively correct account of the consequences of a large magnetic field penetrating the ring. Specifically, we expect that the predictions for the correlation field $B_c$ are parametrically correct while the numerical prefactor would reflect the particular field configuration. At the same time, predictions for the typical current amplitude will be quantitatively correct because the large in-plane field drops out of the final expressions. Some considerations for more general field configurations are collected in an Appendix. \subsection{Toroidal magnetic field} \label{ToroidalMagneticField} The simplification of the toroidal-field model derives from the fact that in this case, the eigenvalue problem \begin{equation} -D[\nabla-{ie}{\bf A}_\pm]^2 \psi = {\cal E}\psi \label{diffuson} \end{equation} together with the boundary condition in Eq.\ (\ref{BoundaryCondition}) can be solved by separation of variables. Let us consider a ring defined as a cylinder of length $L$ (along the $z$-direction) and radius $R$ (in the $x-y$-plane) with periodic boundary conditions in the $z$-direction. The total vector potential ${\bf A}$ is a sum of the Aharonov-Bohm contribution ${\bf A}_\perp= (\phi/L) {\bf \hat z}$ describing the flux threading the ring and the vector potential ${\bf A}_\parallel = (B_\parallel/2) {\bf \hat z}\times{\bf r}$ of the in-plane magnetic field penetrating the ring. Then, the eigenvalue problem in Eq.\ (\ref{diffuson}) separates with $\psi(x,y,z) = \chi(x,y)\exp(ikz)$ where \begin{equation} {\cal E} = E_c (n-\varphi_\pm)^2 + \epsilon_\perp \end{equation} with $n=0,\pm 1,\pm2 \ldots$ and \begin{equation} -D [ (\partial_x -\frac{ieB}{2}y)^2 + (\partial_y +\frac{ieB}{2}x)^2 ] \chi = \epsilon_\perp \chi. \label{DiffusonEquation} \end{equation} Here, we defined the Thouless energy \begin{equation} E_c= \frac{4\pi^2 D}{L^2} \end{equation} and the dimensionless flux variable $\varphi_\pm = \phi_\pm/\phi_0$. Note that in order not to introduce unnecessary numerical prefactors into equations, this definition of the Thouless energy differs by a factor of four from the definitions employed in Refs.\ \onlinecite{Harris09} and \onlinecite{Riedel93}. Inserting these eigenvalues into Eq.\ (\ref{CurrentThroughDifusion}), we find \begin{widetext} \begin{eqnarray} C_I^{(0)}(E,B;E',B') = -\frac{1}{4\pi^2}\frac{\partial^2}{\partial\phi\partial \phi'} \sum_{\pm}\sum_{\epsilon_\perp}\sum_n \int_0^\infty d\sigma {\rm Im} \int_{-2\sigma}^{2\sigma} d\epsilon\, \epsilon \frac{1}{E_c (n-\varphi_\pm)^2 + \epsilon_\perp + i(\epsilon+E-E')}. \end{eqnarray} Performing the sum over $n$ by Poisson summation and measuring all energy variables in units of the Thouless energy, one obtains \begin{eqnarray} C_I^{(0)}(E,B;E',B') = -\frac{E_c^2}{2\pi}\frac{\partial^2}{\partial\phi\partial \phi'} \sum_{\pm}\sum_{\epsilon_\perp}\sum_{p=1}^\infty \cos(2\pi p\varphi_\pm)\int_0^\infty d\sigma {\rm Im} \int_{-2\sigma}^{2\sigma} d\epsilon\, \epsilon \frac{\exp(-2\pi p \sqrt{\epsilon^\pm_\perp + i(\epsilon+E-E')})}{\sqrt{\epsilon^\pm_\perp + i(\epsilon+E-E')}}. \end{eqnarray} The integrals over $\epsilon$ and $\sigma$ can be readily done to yield \begin{eqnarray} C_I^{(0)}(E,B;E',B') &=& -\frac{8E_c^2}{\pi}\frac{\partial^2}{\partial\phi\partial \phi'} \sum_{\pm}\sum_{\epsilon_\perp}\sum_{p=1}^\infty \cos(2\pi p\varphi_\pm) F_p(z_\pm) \label{ZeroTemperature} \end{eqnarray} where $z_\pm=[\epsilon^\pm_\perp + i(E-E')]/E_c$ and where we defined the function \begin{equation} F_p(z) = {\rm Re} \left[ \left( \frac{3}{(2\pi p)^5} + \frac{3\sqrt{z}}{(2\pi p)^4} + \frac{z}{(2\pi p)^3} \right) e^{-2\pi p\sqrt{z}}\right] \label{functioFp} \end{equation} \end{widetext} We note that this result is valid for spinless fermions. Effects of spin will be discussed below in Sec.\ \ref{spin}. It is interesting to compare the result in Eq.\ (\ref{ZeroTemperature}) with the corresponding correlation function for the conductance fluctuations of a metallic ring.\cite{Aronov87,Lee87} Indeed, the flux-sensitive contributions to the correlation function of the conductance at different magnetic fields differ from our result for the persistent current (apart from an overall prefactor) only by the preexponential factor in the function $F_p(z)$. In the absence of the in-plane magnetic field, we need to retain only the lowest transverse eigenvalue $\epsilon_\perp^\pm =0$ to exponential accuracy in $2L/R$. Then, we find \begin{equation} \langle I(\phi)I(\phi')\rangle = \frac{6E_c^2}{\pi^4 \phi_0^2}\sum_{p=1}^\infty \frac{1}{p^3}\sin(2\pi p \varphi)\sin(2\pi p \varphi') \label{IInoSpin} \end{equation} for the current-current correlation, which reproduces the result obtained in Ref.\ \onlinecite{Riedel93}. In the limit of a large in-plane magnetic field, the cooperon contribution is strongly suppressed since time reversal symmetry is broken. This can be seen explicitly by computing the lowest transverse eigenvalue $\epsilon_\perp^\pm$ perturbatively in $B$, for both the cooperon and the diffuson contributions. This perturbative approach is valid as long as $R\ll\ell_B$, where $\ell_B$ has to be evaluated for the appropriate in-plane magnetic fields entering the cooperon ($+$) and diffuson ($-$) contributions. (Here, $\ell_B=(1/eB_\parallel)^{1/2}$ denotes the magnetic length.) Due to the boundary condition of zero normal current, the ground state wavefunction $|{\rm gs}\rangle$ of Eq.\ (\ref{DiffusonEquation}) at zero $B_\parallel$ is a constant with zero transverse eigenvalue. Thus, the leading correction to the eigenvalue is given by \begin{eqnarray} \epsilon_\perp &=& \left\langle {\rm gs}\left| \frac{De^2(B_\parallel)^2}{4}(x^2+y^2)\right|{\rm gs}\right \rangle \nonumber\\ &=& \frac{D}{8\ell_B^2}\left(\frac{R}{\ell_B}\right)^2, \end{eqnarray} and we find that \begin{equation} \frac{\epsilon_\perp}{E_c} = \frac{1}{32\pi^2}\left(\frac{LR}{\ell_B^2}\right)^2. \label{fracEE} \end{equation} For the cooperon contribution, the magnetic field is of the order of twice the applied magnetic field. Thus, by Eq.\ (\ref{ZeroTemperature}) this contribution is exponentially suppressed once the relevant in-plane field is larger than one flux quantum penetrating the ring. We first focus on the typical persistent current at zero temperature. In this case, the effective in-plane field vanishes for the diffuson contribution, while it strongly suppresses the cooperon contribution. Thus, assuming from now on that $B_\parallel$ is sufficiently large to make $\epsilon_\perp(2B)\gg E_c$, we need to retain only the diffuson contribution and obtain \begin{eqnarray} \langle I(\phi)I(\phi')\rangle = \frac{3E_c^2}{\pi^4\phi_0^2}\sum_{p=1}^\infty \frac{1}{p^3} \cos(2\pi p[\varphi-\varphi']). \label{IIhighField} \end{eqnarray} Comparing with Eqs.\ (\ref{HarmonicContent}) and (\ref{RandomPhase}), we find \begin{equation} \langle[I_p^{(+)}]^2\rangle = \langle[I_p^{(-)}]^2\rangle = \frac{1}{2}\langle I_p^2\rangle = \frac{3E_c^2}{\pi^4\phi_0^2}\frac{1}{p^3} \end{equation} for the harmonics of the persistent current. \begin{figure} \includegraphics[width=0.85\columnwidth]{zerotemp.pdf} \caption{(Color online) Current-current correlation function $\langle I(\phi,B_\parallel)I(\phi,B_\parallel+\Delta B_\parallel)\rangle$ (in units of $(E_c/\phi_0)^2$) at zero temperature (solid line). The dashed and dotted lines correspond to the contributions from the first and the second harmonics, respectively. The inset shows the same curves but plotted logarithmically along the vertical axis. \label{CorrelationFunction}} \end{figure} Equation (\ref{ZeroTemperature}), (\ref{functioFp}), and (\ref{fracEE}) also imply that the correlation function of the persistent current at different values of the in-plane magnetic fields falls off exponentially with the magnetic-field difference once the in-plane field changes by more than a flux quantum through the cross section of the ring, i.e., on the scale of the correlation field \begin{equation} B_c = \frac{\sqrt{2}}{\pi}\frac{\phi_0}{LR}. \label{CorrelationField} \end{equation} Note that the functional dependence of the correlation field on $L$ and $R$ remains the same for much more general field configurations but that the numerical prefactor in Eq.\ (\ref{CorrelationField}) is specific to the toroidal-field model. A plot of the correlation function $\langle I(\phi,B_\parallel) I(\phi,B_\parallel + \Delta B_\parallel)\rangle$ is shown in Fig.\ \ref{CorrelationFunction}. Its exponential fall-off has important ramifications in experiment. The decay of the correlation function implies that measurements of the persistent current at in-plane fields which are significantly separated from each other on the scale set by $B_c$ are statistically independent. We are thus led to the ergodic hypothesis that averaging over a sufficiently wide range of in-plane fields is equivalent to averaging over the disorder ensemble. This observation is particularly pertinent in view of the novel technique of measuring persistent currents employed in Ref.\ \onlinecite{Harris09} which allows one to obtain the persistent current over a wide range of in-plane magnetic fields. \begin{figure}[!t] \includegraphics[width=0.85\columnwidth]{finitetemp1.pdf} \caption{(Color online) Current-current correlation function $\langle I(\phi,B_\parallel)I(\phi,B_\parallel+\Delta B_\parallel)\rangle$ (in units of $(E_c/\phi_0)^2$) vs.\ $\Delta B_\parallel$ at finite temperatures. The curves are normalized to their value at $\Delta B_\parallel=0$ and correspond to $T = 0.01, 0.02, 0.1, 0.2, 0.5 \times E_c$ (from bottom to top). The inset shows the same curves but plotted logarithmically along the vertical axis. \label{FiniteTemperaturePC}} \end{figure} We close this section by discussing the temperature dependence of the persistent current at large in-plane magnetic fields. At finite temperatures, the persistent current correlation function depends on $\Delta B$ and temperature $T$ via the two dimensionless variables, $\Delta B/B_c$ and $T/E_c$. The correlation function can be readily evaluated by combining Eq.\ (\ref{FiniteTemperature}) with Eq.\ (\ref{ZeroTemperature}). Performing the remaining integral numerically, we obtain the results shown in Fig.\ \ref{FiniteTemperaturePC} for the current-current correlation function and in Fig.\ \ref{TemperatureDependence} for the temperature dependence of the typical current. We see from Fig.\ \ref{TemperatureDependence} that the temperature dependence can be approximated as exponential with reasonable (though uncontrolled) accuracy. (Numerical values of the fit are quoted in the figure caption.) Moreover, we observe that the typical persistent current becomes rapidly dominated by the first harmonic as temperature increases. \begin{figure}[b] \includegraphics[width=0.85\columnwidth]{TempDep.pdf} \caption{(Color online) Temperature dependence of the typical current $\langle I^2(\phi)\rangle$ (blue line). The dependence can be well fitted by an exponential $\langle I^2\rangle \approx c(E_c/\phi_0)^2\exp(-\alpha T/E_c)$ with $c=0.036$ and $\alpha = 8.2$ as shown by the red (dashed) line. The dotted and dash-dotted lines correspond to the contributions from the first and the second harmonics, respectively. The inset shows the same curves but plotted logarithmically along the vertical axis. \label{TemperatureDependence}} \end{figure} \subsection{Effects of spin} \label{spin} In weak magnetic field and in the absence of spin-orbit scattering, spin enters the persistent current simply through a degeneracy factor of two. Thus, Eq.\ (\ref{IInoSpin}) is modified into \begin{equation} \langle I(\phi)I(\phi')\rangle = \frac{24E_c^2}{\pi^4 \phi_0^2}\sum_{p=1}^\infty \frac{1}{p^3}\sin(2\pi p \varphi)\sin(2\pi p \varphi'). \end{equation} This result includes both the diffuson and cooperon contributions. In a large applied magnetic field, but still without spin-orbit scattering, the cooperon contribution is suppressed and we have to take the Zeeman energy into account. The corresponding spinless result was given in Eq.\ (\ref{IIhighField}). We can include the spin and Zeeman energies by writing the persistent current as a sum of the contributions of spin-up and spin-down electrons, $I=I_\uparrow+I_\downarrow$. Once the Zeeman energy becomes large compared to the Thouless energy, there are no correlations between $I_\uparrow$ and $I_\downarrow$ and as a result, we find \begin{equation} \langle I(\phi)I(\phi')\rangle = \frac{6E_c^2}{\pi^4 \phi_0^2}\sum_{p=1}^\infty \frac{1}{p^3}\cos(2\pi p [\varphi- \varphi']). \label{ZeroTemperatureII} \end{equation} The recent precision measurements\cite{Harris09} of the persistent current were performed on samples whose spin-orbit scattering length is smaller than or of order of the circumference of the rings, as deduced from weak-localization measurements. For this reason, we now turn to a more thorough discussion of the consequences of the electron spin, which in addition accounts for the spin-orbit scattering. This can be done by a standard extension of the diagrammatic technique for diffusive systems.\cite{Vavilov03} To be specific, we focus on sufficiently large magnetic fields that the cooperon no longer contributes significantly. Extensions to include the cooperon contribution at weak fields would pose no additional complications. Including spin indices, we define the diffuson ${\cal D}_{s^\prime_1 s^\prime_2}^{s_1 s_2}({\bf r},{\bf r}', \epsilon)$ as shown in Fig.\ \ref{DiffusonDiagram} and view it as a $4\times4$-matrix ${\bf D}({\bf r},{\bf r}', \epsilon)$ where $(s_1, s_1^\prime)$ labels the rows and $(s_2, s_2^\prime)$ the columns. With the ordering $(s,s')=(\uparrow\uparrow, \uparrow\downarrow, \downarrow\uparrow, \downarrow\downarrow)$, one obtains the equation\cite{Vavilov03} \begin{widetext} \begin{equation} \left[-D\left(\nabla-{ie}{\bf A_-}\right)^2 +i\epsilon +H_Z+H_{\rm so}\right]{\bf D}({\bf r},{\bf r}', \epsilon) = \frac{1}{2\pi N(0)\tau^2} \delta({\bf r}-{\bf r}') \end{equation} by the standard procedure, starting with the diagrammatic representation shown in Fig.\ \ref{DiffusonDiagram}. ($N(0)$ denotes the density of states at the Fermi energy and $\tau$ is the elastic scattering time.) Here, the contribution of the Zeeman energy $E_Z$ yields the term \begin{equation} H_Z = \left[\begin{array}{cccc} 0 &0 & 0 & 0 \\ 0 & -2iE_Z & 0 & 0 \\ 0 & 0 & 2iE_Z & 0 \\ 0 & 0 & 0 & 0 \end{array}\right], \end{equation} while spin-orbit scattering is included through \begin{equation} H_{\rm so} = \frac{2}{3\tau_{\rm so}}\left[\begin{array}{cccc} 1 &0 & 0 & -1 \\ 0 & 2 & 0 & 0 \\ 0 & 0 & 2 & 0 \\ -1 & 0 & 0 & 1 \end{array}\right] \end{equation} in terms of the spin-orbit scattering time $\tau_{\rm so}$. By retracing the steps leading up to Eq.\ (\ref{CurrentThroughDifusion}) in the presence of spin effects, we obtain for the correlation function of the persistent current, \begin{eqnarray} C_I^{(0)}(E,B;E',B') = -\frac{1}{4\pi^2}\frac{\partial^2}{\partial\phi\partial \phi'} \int_0^\infty d\sigma {\rm Im} \int_{-2\sigma}^{2\sigma} d\epsilon\, \epsilon {\rm Tr}\left(\frac{1}{-D[\nabla-{ie}{\bf A}_-]^2 + i(\epsilon+E-E') + H_Z + H_{\rm so}}\right), \label{CurrentThroughDifusionWithSpin} \end{eqnarray} where ${\rm Tr}$ now denotes a trace over configuration space and the four-dimensional spin space. In the limit of large Zeeman splitting, $E_Z \gg E_c$, the modes $\uparrow\downarrow$ and $\downarrow\uparrow$ are exponentially suppressed. For negligible spin-orbit scattering, we then obtain two massless modes $\uparrow\uparrow\pm \downarrow\downarrow$. As a result, the correlation function is twice larger than the result for spinless electrons given in Eq.\ (\ref{ZeroTemperature}), in agreement with Eq.\ (\ref{ZeroTemperatureII}). As the spin-orbit scattering increases, only the density mode $\uparrow\uparrow + \downarrow\downarrow$ remains massless and in the limit of strong spin-orbit scattering, we recover the result in Eq.\ (\ref{ZeroTemperature}) for spinless electrons. More generally, we can discuss the crossover between the limits of weak and strong spin-orbit scattering rate. One finds \begin{eqnarray} &&C_I^{(0)}(E,B;E',B') = -\frac{1}{4\pi^2}\frac{\partial^2}{\partial\phi\partial \phi'} \int_0^\infty d\sigma {\rm Im} \int_{-2\sigma}^{2\sigma} d\epsilon\, \epsilon \nonumber\\ && \,\,\,\,\,\,\,\, \times{\rm Tr}\left(\frac{1}{-D[\nabla-{ie}{\bf A}_-]^2 + i(\epsilon+E-E')} + \frac{1}{-D[\nabla-{ie}{\bf A}_-]^2 + i(\epsilon+E-E') + \frac{4}{3\tau_{\rm so}}}\right. \nonumber\\ && \,\,\,\,\,\,\,\, + \left.\frac{1}{-D[\nabla-{ie}{\bf A}_-]^2 + i(\epsilon+E-E'+2E_Z) + \frac{4}{3\tau_{\rm so}}} +\frac{1}{-D[\nabla-{ie}{\bf A}_-]^2 + i(\epsilon+E-E'-2E_Z) + \frac{4}{3\tau_{\rm so}}}\right) , \end{eqnarray} where the trace is now over configuration space only. Specifying again to the toroidal-field model, we obtain \begin{eqnarray} &&C_I^{(0)}(E,B;E',B') = -\frac{8E_c^2}{\pi}\frac{\partial^2}{\partial\phi\partial \phi'}\sum_{p=1}^\infty \cos(2\pi p\varphi_\pm) \left[F_p\left(i\frac{E-E'}{E_c}+ \frac{\epsilon^-_\perp}{E_c}\right) \right. \nonumber\\ &&\, \left. + F_p\left(i\frac{E-E'}{E_c}+\frac{\epsilon^-_\perp+ \frac{4}{3\tau_{\rm so}}}{E_c}\right) + F_p\left(i\frac{E-E'+2E_Z}{E_c}+\frac{\epsilon^-_\perp+ \frac{4}{3\tau_{\rm so}}}{E_c}\right) + F_p\left(i\frac{E-E'-2E_Z}{E_c}+\frac{\epsilon^-_\perp+ \frac{4}{3\tau_{\rm so}}}{E_c}\right) \right] \label{cross_over} \end{eqnarray} \end{widetext} where the function $F_p(z)$ had been defined in Eq.\ (\ref{functioFp}). \begin{figure}[b] \includegraphics[width=0.85\columnwidth]{diffuson.pdf} \caption{Diagrammatic representation of the equation of motion for the diffuson ${\cal D}_{s^\prime_1 s^\prime_2}^{s_1 s_2}({\bf r},{\bf r}', \epsilon)$. Full lines represent electronic Green functions and dashed lines denote disorder and spin-orbit scattering. \label{DiffusonDiagram}} \end{figure} Combining Eq.\ (\ref{cross_over}) with Eq.\ (\ref{FiniteTemperature}) and setting $\epsilon^-_\perp=0$, we can obtain the crossover of the typical current between the limits of weak and strong spin-orbit scattering for arbitrary temperature. (Note that the results for the typical current are not restricted to the toroidal-field model.) Corresponding numerical results in the limit of large Zeeman splitting (where the last two terms in the square bracket in Eq.\ (\ref{cross_over}) can be neglected) are plotted in Fig.\ \ref{crossover}, which show that the crossover becomes slower as temperature increases. \begin{figure}[b] \includegraphics[width=0.85\columnwidth]{spinorbit.pdf} \caption{(Color online) Crossover of the typical current $\langle I^2(\phi)\rangle$ as function of the spin-orbit scattering rate. The curves, corresponding to temperatures $T = 0.01, 0.1, 0.3, 1.0 \times E_c$ (from bottom to top), are normalized to the value of $\langle I^2(\phi)\rangle$ in the limit of vanishing spin-orbit scattering rate. All curves are plotted in the limit of large in-plane field where the cooperon contribution is suppressed and the Zeeman energy is large compared to the Thouless energy. \label{crossover}} \end{figure} \section{Interaction contribution} \label{interaction} We now turn to a discussion of the interaction-contribution to the persistent current in high magnetic fields. Adapting the first-order correction in the interaction $V$ derived in Ref.\ \onlinecite{Ambegaokar90} to the case of a finite magnetic field, one finds for the disorder-averaged contribution to the grand canonical potential \begin{eqnarray} \label{IntGrandPot-1} \Delta \Omega &=& \frac{N(0)\bar{V}}{\pi}\int_0^{\infty} dE \coth (\frac{E}{2T})E \nonumber \\ &&\times {\rm Re}{\rm Tr} \frac{1}{-D(\nabla-{2ie}{\bf A})^2+iE} \end{eqnarray} Here, $\bar{V}$ is the Fourier component of the screened Coulomb interaction potential averaged in momentum space.\cite{Ambegaokar90} In a field much stronger than the upper critical field of the ring we may constrain considerations to the lowest-order correction, Eq.~(\ref{IntGrandPot-1}). To estimate the interaction contribution to the average persistent current, we again employ the toroidal-field model introduced in Sec.\ \ref{independent}. Then, the eigenvalue problem and boundary conditions for the cooperon here are identical to those in Eqs.\ (\ref{BoundaryCondition}) and (\ref{diffuson}), respectively. We denote the cooperon eigenvalues by $\epsilon^{(l)}_{n,m,\phi}$ with $l,n,m$ being the radial, longitudinal and azimuthal quantum numbers, respectively. Due to cylindrical symmetry the cooperon modes can be found by separation of variables, with the replacement $n \rightarrow n-2\phi/\phi_0$ added to take into account the Aharonov-Bohm flux. In distinction from Sec.\ \ref{independent}, the vector potential ${\bf A}$ in Eq.\ (\ref{IntGrandPot-1}) corresponds to the total field so that $\ell_B \ll R$. In this limit, the radial equation can be approximated to lowest order in $\ell_B/R$ as \begin{equation}\label{IntEigenEq} \frac{D}{\ell_B^2}\left( -\frac{\partial^2}{\partial x^2} + (\kappa_m-x)^2+\ell_B^2 k_{n,\phi}^2\right)\chi(x)=\epsilon^{(l)}_{n,m,\phi}\chi(x) \end{equation} where $x=r/\ell_B$ is a scaled distance from the center of the cross section, $\kappa_m=m \ell_B/R$ and $k_{n,\phi}=2\pi(n-2\phi/\phi_0)/L$. Note that the ratio between the radial and the longitudinal terms in Eq.\ (\ref{IntEigenEq}) is dominated by $L/\ell_B$. The eigenvalues can be written as \begin{equation} \epsilon^{(l)}_{n,m,\phi}=D\left(\frac{2\pi}{L}\right)^2\left[ (n-\frac{\phi}{\phi_0/2})^2+\left(\frac{L}{2\pi \ell_B}\right)^2\lambda_l(\kappa_m) \right] \end{equation} where the values of $\lambda_0(\kappa_m)$ for the lowest branch of eigenstates ($l=0$) can be estimated by using the variational method with a Gaussian trial solution. The function $\lambda_0(\kappa)$ has a shallow minimum $\lambda_0^*=(1-2/\pi)^{1/2}$ at $\kappa_m^*=(\pi^2-2\pi)^{-1/4}$. Using the eigenvalues $\epsilon^{(l)}_{n,m,\phi}$ to evaluate the trace in Eq.~(\ref{IntGrandPot-1}), it is straightforward to show that the contribution to persistent current $\Delta I=-\partial \Delta \Omega /\partial \phi$ is periodic in $\phi\rightarrow \phi+\phi_0/2$, and for $T=0$ can be written as \begin{equation} \Delta I=\frac{N(0)\bar{V}}{\pi} \left( \frac{2\pi}{L}\right)^2\frac{4hD}{\phi_0}\sum_{p=1}^{\infty} p g_p \sin\left( 2\pi p \frac{\phi}{\phi_0/2}\right)\,. \end{equation} In the regime of experimental interest, $L\gg R \gg \ell_B$, the coefficients $g_p$, \begin{equation} g_p=\frac{1}{2p^3 \pi^2}\sum_{m=0}^{\infty} e^{-pL/\ell_B \sqrt{\lambda_0(\kappa_m)}}\left(1+ p \frac{L}{\ell_B}\sqrt{\lambda_0(\kappa_m)}\right) \end{equation} can be estimated by evaluating the sum in the saddle-point approximation, \begin{equation} g_p\approx 0.13p^{-3.5}\left( \frac{R}{\sqrt{\ell_B L}}\right) \left[1+p\frac{L}{\ell_B}\sqrt{\lambda_0^*}\right]e^{-\frac{L}{\ell_B}p\sqrt{\lambda_0^*}}\,. \end{equation} All harmonics of the average persistent current are exponentially suppressed; the higher the harmonic $p$, the stronger is the suppression. Note that this implies that for sufficiently strong magnetic field, measurements of the average current, e.g., by employing large arrays of rings, should be dominated by the canonical-ensemble contribution of the free-electron model.\cite{Altshuler91, Schmid91, Oppen91} \begin{figure}[t] \includegraphics[width=0.85\columnwidth]{IvsTdatafit.pdf} \caption{(Color online) Temperature dependence of the typical current contribution from the $p^{\rm th}$ harmonic $\sqrt{\langle I_p^2(T) \rangle}$. The markers are the data first presented in Ref.\ \onlinecite{Harris09}. The solid curves represent new fits to the data using Eqs.\ (\ref{FiniteTemperature}) and (\ref{cross_over}) while the dashed curves show the fits from Ref.\ \onlinecite{Harris09}. The sample parameters and best-fit parameters are given in Table \ref{diffusionconstants}. Closed and open markers denote measurements taken during different cooldowns, over different field ranges, and at different magnetic field orientations. In the case of the $p=1$ data from Sample \#1, the two different field ranges over which the closed and open markers were taken lead to slightly different values of the fitting function at high temperature, with the lower curve corresponding to the closed markers and the upper curve to the open markers. The new fit curves for Samples \#2 \& \#3 are indistinguishable, as are the old and new fit curves for Sample \#4. \label{datafit}} \end{figure} \section{Comparison to Experiment} \label{datacomparison} The recent development of cantilever-based torsional magnetometers with integrated mesoscopic rings\cite{Bleszynski08} resulted in measurements of the rings' persistent current in the presence of large magnetic fields.\cite{Harris09} Here we briefly review these measurements and compare them with the calculations from the preceding sections. This comparison is most readily performed by fitting the measured temperature dependence of the current to the form predicted in Eqs.\ (\ref{FiniteTemperature}) and (\ref{cross_over}). \begin{table*} \caption{Sample parameters. ``Marker'' refers to the markers used in Fig.\ \ref{datafit}, with closed and open markers representing two different cooldowns of the same sample. For the closed markers the angle between the magnetic field and the plane of the rings was 6$^{\circ}$ and $T_0=323$ mK, while for the open markers the angle was 45$^{\circ}$ and $T_0=365$ mK. $N$ denotes the number of rings in the sample. The ring circumference and linewidth are given by $L$ and $w$. The thickness of each sample was 90 nm. The spin orbit scattering length $L_{\rm so}=1.1\: \pm\: 0.25 \:\mu \mathrm{m}$. $B_{\rm min}$ and $B_{\rm max}$ give the bounds for measurements of $I(B)$ taken over smaller field ranges. $D_L$ and $D_{\rm ZSO}$ are extracted from fitting the persistent current data. $D_L$ is the best-fit value of the diffusion constant found in Ref.\ \onlinecite{Harris09}, which assumed the limit of strong spin-orbit scattering and large Zeeman splitting. $D_{\rm ZSO}$ is the best-fit value of the diffusion constant found by taking into account the finite spin-orbit scattering rate and Zeeman splitting as described in Section\ \ref{datacomparison}. The estimated uncertainty in all fit coefficients is 6\%. } \begin{tabular*}{1.7\columnwidth}{@{\extracolsep{\fill}} | c c c c c c c c c c |}\hline Sample & Marker & $p$ & $N$ & $L$ ($\mu$m) & $w$ (nm) & $B_{\rm min}$ (T) & $B_{\rm max}$ (T) & $D_{L}$ (cm$^2$/s) & $D_{\rm ZSO}$ (cm$^2$/s) \\ \hline \multirow{3}{*}{\#1} & \includegraphics{CL17_6D.pdf} & 1 & \multirow{3}{*}{1680} & \multirow{3}{*}{1.9} & \multirow{3}{*}{115} & 6.2 & 6.8 & \multirow{3}{*}{271} & \multirow{3}{*}{234} \\ & \includegraphics{CL17H1_45D.pdf} & 1 & & & & 5.0 & 5.2 & & \\ & \includegraphics{CL17H2_45D.pdf} & 2 & & & & 5.0 & 5.2 & & \\ \hline \multirow{2}{*}{\#2} & \includegraphics{CL15_6D.pdf} & \multirow{2}{*}{1} & \multirow{2}{*}{990} & \multirow{2}{*}{2.6} & \multirow{2}{*}{85} & 7.15 & 7.60 & \multirow{2}{*}{214} & \multirow{2}{*}{195} \\ & \includegraphics{CL15_45D.pdf} & & & & & 5.39 & 5.48 & & \\ \hline \#3 & \includegraphics{CL14_45D.pdf} & 1 & 1 & 2.6 & 85 & 8.32 & 8.40 & 215 & 195 \\ \hline \#4 & \includegraphics{CL11_6D.pdf} & 1 & 242 & 5.0 & 85 & 7.1 & 7.3 & 205 & 196 \\ \hline \end{tabular*} \label{diffusionconstants} \end{table*} The parameters characterizing each sample are collected in Table\ \ref{diffusionconstants}. The temperature dependence of the $p^{\rm th}$ harmonic $\sqrt{\langle I_p^2\rangle}$ of each sample's typical current was determined as follows. At a single temperature $T_0$, the mean square amplitude of the $p^{\rm th}$ harmonic of the current was extracted from a measurement of $I(B)$ taken over a range of $B$ spanning many $B_c$. This large span ensured that the mean was determined from a large number of independent measurements, as discussed at the end of Sec.\ \ref{ToroidalMagneticField}. For each sample, the form of $I(B)$ was found to be independent of temperature except for an overall scaling. This scaling was determined by measuring $I(B)$ over a smaller field range (with bounds denoted by $B_{\rm min}$ and $B_{\rm max}$) at each subsequent temperature and comparing the magnitude of each harmonic with the value measured over the same field range at $T_0$. This procedure, as well as other details of the measurements, are described in detail in Ref.\ \onlinecite{Harris09}. The resulting values of $\sqrt{\langle I_p^2\rangle}$ are shown in Fig.\ \ref{datafit}. In Ref.\ \onlinecite{Harris09}, this data was analyzed by assuming the limit of strong spin orbit scattering: $1/\tau_{\rm so}\gg \{E_c, T\}$, and large Zeeman splitting, $E_Z\gg \{E_c,T\}$. As can be seen from the sample parameters listed in Table\ \ref{diffusionconstants}, this assumption is fairly accurate though not exact. For these samples $0.075 < 1/E_c \tau_{\rm so} < 0.47$ while $0.15 < T/E_c < 1.7$. From Fig.\ \ref{crossover} it is clear that these parameters are not fully within the strong spin orbit scattering limit. Additionally, for the smallest rings (Sample \#1) the limit of large Zeeman splitting does not hold at the highest temperatures in Fig.\ \ref{datafit} where $ T/E_Z \approx 0.36$ and deviations from the large Zeeman splitting limit change $\sqrt{\langle I_p^2\rangle}$ by as much as 5\%. For Samples \#2, \#3, and \#4 $ T/E_Z < 0.17$ resulting in $< 1\%$ deviations of $\sqrt{\langle I_p^2\rangle}$ from the large Zeeman splitting limit. As a result the measurements of Ref.\ \onlinecite{Harris09} were not fully in the strong spin orbit scattering, large Zeeman splitting limit, so we reanalyze the data here, taking into account the full dependence of $\sqrt{\langle I_p^2\rangle}$ on spin orbit scattering and Zeeman splitting. We fit the data from Ref.\ \onlinecite{Harris09} (Fig.\ \ref{datafit}) using the expression for $\sqrt{\langle I_p^2 (T,D,L_{\rm so},E_z)\rangle}$ derived from Eqs.\ (\ref{FiniteTemperature}) and (\ref{cross_over}). The only fitting parameter is the electron diffusion constant $D$. The spin orbit length $L_{\rm so}\equiv\sqrt{D \tau_{\rm so}} = 1.1\: \pm\: 0.25\: \mu \mathrm{m}$ was determined independently from magnetotransport measurements of a wire codeposited with the rings.\cite{Harris09} Since each data point in Fig.\ \ref{datafit} is extracted from measurements of $I(B)$ made over a range of $B$, we cannot use a single value of the Zeeman splitting; instead, we average over the magnetic field range to obtain the fitting function \begin{eqnarray} \lefteqn{\sqrt{\langle I_p^2 (T,D,L_{\rm so},B_{\rm min},B_{\rm max})\rangle} = } \nonumber \\ & & \sqrt{\frac{\int_{B_{\rm min}}^{B_{\rm max}}dB\langle I_p^2 (T,D,L_{\rm so},E_Z(B))\rangle}{B_{\rm max}-B_{\rm min}}}. \label{EZaverage} \end{eqnarray} The best-fit values of the diffusion constant $D_{\rm ZSO}$ are given in Table\ \ref{diffusionconstants}. The corresponding fits are shown in Fig.\ \ref{datafit} as solid lines. For comparison the values of the diffusion constant found in Ref.\ \onlinecite{Harris09}, $D_L$, are also given in Table\ \ref{diffusionconstants} and the corresponding fits are shown as dashed lines in Fig.\ \ref{datafit}. Figure\ \ref{datafit} and Table\ \ref{diffusionconstants} show that the finiteness of the spin-orbit scattering rate and the Zeeman energy result in small but noticeable changes to the fitted curves and the extracted values of $D$. We find that most of the difference is due to the finite spin orbit scattering rate, which leads to a non-negligible contribution to the current from the second $F_p$ term in Eq.\ (\ref{cross_over}). The finite Zeeman energy modifies the current via the last two $F_p$ terms in Eq.\ (\ref{cross_over}), leading to a correction which becomes appreciable ($> 1 \%$) only for the higher temperature measurements of Sample \#1. The resulting correction oscillates as a function of temperature, resulting in a best-fit value of $D$ indistinguishable from the case of large Zeeman splitting. The values of $D_{\rm ZSO}$ for Samples \#2, \#3, and \#4 agree with each other to within the experimental uncertainty (which is estimated to be $6 \%$ in the Supplemental Online Material of Ref.\ \onlinecite{Harris09}). This agreement is consistent with the fact that the rings in these three samples have the same cross-sectional dimensions. The value of $D_{\rm ZSO} = 234 \:\mathrm{ cm}^2/\mathrm{s}$ measured for Sample \#1 is somewhat larger, which may reflect these rings' larger cross section. Resistivity measurements of the codeposited wire having the same cross section as Sample \#1 give $D = 260 \pm 12 \:\mathrm{ cm}^2/\mathrm{s}$, consistent with the value measured for Sample \#1.\cite{Harris09} \section{Conclusions} \label{conclusions} Motivated by a new and highly sensitive experimental technique \cite{Harris09} for measuring mesoscopic persistent currents, we presented a theory of persistent currents in large, but non-quantizing, magnetic fields. The theoretical results of this paper formed the basis for establishing the remarkable quantitative agreement between experiment and theory found in Ref.\ \onlinecite{Harris09} and further refined in Sec.\ref{datacomparison}. To reach this agreement, we not only needed to take into account the large magnetic field, both for the single-particle and the interaction contributions to the persistent current, but also spin effects. In addition to forming the basis for a quantitative comparison with experiment, it is also worth emphasizing several theoretical conclusions from our results. (i) The magnetic field penetrating the ring leads to qualitative changes in the dependence of the persistent current on the Aharonov-Bohm flux. At zero magnetic field, the persistent current is a periodic function of flux. Zero flux as well as integer and half-integer multiples of the flux quantum are special points where the persistent current vanishes. At large magnetic fields, the persistent current $I(\phi)$ is still a periodic function of flux, but the typical magnitude $\langle I^2(\phi)\rangle $ is no longer dependent on flux. (ii) Previous theoretical works have shown that there are two principal contributions to mesoscopic persistent currents: a free-electron contribution and an interaction contribution. In experiments, it is not always easy to disentangle these two contributions (especially for the even harmonics of the persistent current). In fact, while the interaction contribution is expected to dominate the ensemble-averaged persistent current, both of them contribute significantly in single- or few-ring experiments. We conclude from our results that the application of a large magnetic field penetrating the ring strongly suppresses the interaction contribution to the persistent current so that the technique of Ref.\ \onlinecite{Harris09} provides direct access to the free-electron contribution. (iii) One of the principal advantages of the experimental technique of Ref.\ \onlinecite{Harris09} is that unlike SQUID-based approaches, it allows for measurements over a wide range of magnetic fields and thus of many oscillations of the persistent current with flux. Our results for the autocorrelation function of the persistent current at different magnetic fields imply that averaging over magnetic field is equivalent to an ensemble average (ergodic hypothesis). One of the possibilities raised by this result is a direct measurement of the entire distribution function of the persistent current. The experimental technique of Ref.\ \onlinecite{Harris09} has brought many additional experiments on persistent currents and related phenomena within experimental reach. Our approach should be a valuable starting point for analyzing such future experiments. \begin{acknowledgments} This work was supported in part by DOE grant DE-FG02-08ER46482 (LG), by DIP (FvO), as well as by NSF grants 0706380 and 0653377 (JGEH). FvO and LG acknowledge the hospitality of KITP while part of this work has been performed. \end{acknowledgments}
\section{Introduction} Dwarf galaxies are thought to dominate the cosmic scenery in terms of number density \citep[][]{springel_2005,sandage_virgo_LF,mateo_dwarfs_review_1998} and therefore form an important morphological class of galaxies. Several investigators have studied the mass distributions of these galaxies \citep[e.g.][and references therein]{carignan_beaulieu_DDO154_1989,cote_carignan_sancisi_1991,broeils_phd,meurer2,deblok_mcgaugh_1996}. \citet{swaters_thesis} was the first to study the dark matter (DM) properties of a large representative sample of nearby dwarf galaxies as part of the Westerbork H\,{\sc i}\ Survey of Spiral and Irregular Galaxies \citep[WHISP, ][]{WHISP_swaters}. More recently, a morphologically diverse sample of nearby galaxies, including dwarf systems, was observed as part of The~H\,{\sc i}\ Nearby~Galaxy~Survey \citep[THINGS, ][]{THINGS_walter}. These cumulative efforts have led to an understanding that DM plays an important role in the dynamics and evolution of dwarf galaxies. \citet{swaters_thesis} found that none of the observed rotation curves in his sample of late-type dwarf galaxies decline at outer radii. This was further confirmation that these systems have a large dark mass fraction. Dwarf galaxies therefore serve as ideal candidates to test theories of DM. Observational studies of the mass distributions of these dwarf systems have shown their DM halos to be well-modelled by a pseudo iso-thermal sphere with an approximately constant-density core \citep{deblok_et_al_1996,cote_carignan_freeman_2000,THINGS_deblok}. Theoretically, the study of the properties of DM halos in a hierarchical clustering context is carried out mainly in the form of Cold Dark Matter (CDM) numerical simulations \citep[e.g.][]{NFW,moore_1999,aquarius_subhalos}. The equilibrium density profile, $\rho(r)$, of simulated DM halos is found to vary with radius, $r$, as $\rho(r) \propto r^{\alpha}$ with $\alpha \approx -1$ \citep{NFW_1997}, thereby predicting extremely steep inner density profiles. This discrepancy between the observed shapes of DM halos and the predictions from numerical simulations has become known as the ``cusp/core'' problem. Attempts at reconciling observations with theoretical predictions depend crucially on accurate dynamical analyses of nearby galaxies. To test DM halo models one wants galaxies that are as dark-matter-dominated as possible. In this respect, a potentially useful candidate for DM studies is NGC~2915. This galaxy is classified as a nearby \citep[$D\sim$~3.78~Mpc, ][]{karachentsev_catalog} blue compact dwarf according to its optical appearance, yet has the H\,{\sc i}\ morphology of a late-type spiral. Detailed imaging by \citet{meurer1} showed that the optical appearance of NGC~2915 is dominated by two main stellar populations: a compact blue population, which is the location of on-going high-mass star formation; and a more diffuse, older red population. What makes this galaxy so interesting is the fact that, when observed at 21~cm, the stellar core of NGC~2915 is seen to be completely embedded in a huge H\,{\sc i}\ disk extending out to $R\sim$~22 $B$-band scale lengths \citep{meurer2}. Furthermore, this gas disk has well-defined spiral structure that is not seen in the stellar disk. No significant star-formation is observed in the outer parts of this gas disk. Only a few faint HII regions were detected by \citet{meurer_1999} after carrying out deep H$\alpha$ imaging of the disk. The H\,{\sc i}\ disk of NGC~2915 serves as an ideal tracer of the gravitational potential out to radii of $R\sim$~10~kpc, far beyond the visible radial extent of the stellar disk. Due to the apparent lack of stars in its outer parts, the dynamics of the H\,{\sc i}\ disk are an almost direct tracer of the DM distribution. \citet{meurer2} targeted NGC~2915 for DM studies and found the galaxy to be DM-dominated at nearly all radii with a total mass to $B$-band light ratio M$_{tot}$/L$_B$~$\gtrsim$~76~$M_{\odot}$/L$_{\odot}$ (assuming a distance of 5.1 Mpc). They also found the galaxy to have a dense ($\rho_0~\approx~0.1$~$M_{\odot}$~pc$^{-3}$) and compact ($r_c\approx$~1~kpc) DM core. Several investigators have attempted to explain the observed H\,{\sc i}\ distribution of NGC~2915. \citet{bureau_1999} studied the dynamics of the central H\,{\sc i}\ region and the spiral pattern of the outer disk. For both they calculated a common, slow pattern speed of $\Omega_p=8.0\pm 2.4$~km~s$^{-1}$\ which they associated with the figure rotation of a tri-axial DM halo. They also proposed that some DM is distributed in the disk of NGC~2915, thereby making it gravitationally unstable to the formation of the observed spiral structure. \citet{masset_bureau_2003}, using hydrodynamical simulations, further explored the ideas of \citet{bureau_1999}. They showed that the observed spiral structure can be accounted for by either an unseen bar or a rotating tri-axial DM halo. However, the mass of the required bar, $M_{bar}\sim 5\times 10^9$~$M_{\odot}$, is very large in comparison to the total stellar mass, thereby making the nature of such a bar problematic. The required pattern speed of the tri-axial DM halo is significantly larger than those from numerical simulations. \citet{masset_bureau_2003} disfavoured the external perturber scenario. They also found that while a heavy disk is able to account for the main features of the observed H\,{\sc i}\ morphology, it fails to match the observed gas dynamics. A satisfactory explanation for the various morphological and kinematic features of the H\,{\sc i}\ in NGC~2915 is therefore still lacking. In this paper we investigate the H\,{\sc i}\ dynamics and DM distribution of NGC~2915. We use data from new H\,{\sc i}\ synthesis observations of NGC~2915 carried out using the Australian Telescope Compact Array as part of the Southern Hemisphere extension of The H\,{\sc i}\ Nearby Galaxy Survey. Our H\,{\sc i}\ observations are significantly deeper and have better spatial resolution than any other H\,{\sc i}\ observations of NGC~2915. These data are complemented by high-quality 3.6~\micron\ infrared observations of the stellar disk, carried out as part of the \emph{Spitzer} Infrared Nearby Galaxies Survey \citep[SINGS,][]{SINGS}. The H\,{\sc i}\ observations are presented in Sec.~\ref{data_cube} while the H\,{\sc i}\ data products appear in Sec.~\ref{data_products}. This paper focuses on the regular gas dynamics of NGC~2915. In Sec.~\ref{vrot} a rotation curve out to $R\sim$~9.3~kpc is determined. The far-infrared observations of the stellar disk are combined with our new H\,{\sc i}\ observations in Sec.~\ref{mass_modeling} to produce a mass model that determines the contributions at various radii of the stars, gas and DM. We use two different parameterisations of the DM halo to reconstruct the observed rotation curve, one of which is the NFW halo \citep{NFW_1997} favoured by numerical simulations while the other is the observationally motivated pseudo-isothermal sphere. Finally, in Sec.~\ref{conclusions}, we summarise our results and present our conclusions. \section{H\,{\sc i}\ observations and data reduction}\label{data_cube} \subsection{Data acquisition} NGC 2915 was observed between 23~October~2006 and 2~June~2007 (project number C~1629) with six different ATCA configurations using all six antennas. A single run consisted of a primary calibrator observation, regular secondary calibrator observations and source observations. The calibrator sources PKS~1934--63 and PKS~0823--500 were used as primary and secondary calibrators respectively. Each of the EW-352, 750D, 1.5B and 1.5C runs was approximately 12 hours long while 24 hours were spent in the 6A configuration. Besides these data, archival data were also incorporated\footnote{Project number C191, principle investigator: Meurer.}. \citet{meurer2} determined the H\,{\sc i} diameter of NGC 2915 to be $\sim$~0.32$^{\circ}$. The H\,{\sc i} extent of NGC 2915 therefore falls well within the $\sim$~0.54$^{\circ}$ field-of-view of the ATCA dishes when observing at 21~cm. For our observations, the telescope pointing centres were set to the optical centre of NGC 2915 ($\alpha_{2000}$~=~09$^{\mathrm{h}}$~26$^{\mathrm{m}}$~11.5$^{\mathrm{s}}$, $\delta_{2000}$~=~-76$^{\circ}$~37$'$~35$''$) with no mosaicking required. The correlator was set to use 512 channels with a bandwidth of 8~MHz, centered at 1418~MHz. The resulting velocity range is $-335$~km~s$^{-1}$\ to 1353~km~s$^{-1}$\ with an approximate channel spacing of 3.2~km~s$^{-1}$. Table \ref{ATCA_table} provides a summary of all our observing setups. \begin{table} \begin{center} \caption{Summary of NGC 2915 observing setups} \label{ATCA_table} \scriptsize{ \begin{tabular}{ccccccc} \hline \hline \\ 1 & 2 & 3 & 4 & 5\\ Conf. & Date & Start & End & Dur. \\ × & (yy-mm-dd) & (hh-mm-ss) & (hh-mm-ss) & (hrs) \\ \\ \hline \\ EW352 & 2006-10-23 & 11:28:55 & 23:17:45 & 10.39 \\ \\ 750D & 2007-03-14 & 04:58:45 & 15:56:10 & 9.64 \\ \\ 1.5B & 2006-11-24 & 19:29:05 & 06:51:35 & 9.95 \\ \\ 1.5C & 2007-05-02 & 04:37:25 & 15:36:05 & 9.64 \\ \\ 6A & 2007-02-10 & 10:34:35 & 21:32:25 & 9.64 \\ \\ 6A & 2007-02-16 & 05:12:35 & 00:42:15& 17.06 \\ \\ \hline \end{tabular}} \end{center} \textbf{Comments on columns:} Column 1: ATCA configuration used; Column 2: date of observation (UT); Column 3/4: start/end of observations (UT); Column 5: time on source. \end{table} \subsection{H\,{\sc i} Data cubes} The MIRIAD software package \citep{MIRIAD} was used to take the raw $uv$ data from the correlator through to the image analysis stage. All velocities are measured relative to the barycentre rest frame. The first five and last five correlator channels were flagged. The data were then split into primary calibrator, secondary calibrator, and source subsets. The required corrections to the antenna gains, delay terms and bandpass shapes were determined. The time-varying phases and antenna gains were calibrated based on observations of the secondary calibrator. The calibrated source data were continuum-subtracted by fitting and subtracting a first-order polynomial to the line-free channels which, for our observations, were channels $50-100$ and $350-460$. Having isolated the H\,{\sc i} signal, image cubes were produced from the continuum-subtracted $uv$ data. Data cubes were produced using natural weighting as well as robust weighting with a Brigg's visibility weighting robustness parameter of 0.2. Each of the dirty images was deconvolved using a Steer Clean algorithm \citep{steer_clean} to produce an output map of clean components. Channels were either cleaned down to a flux cut-off of 2.5 times the typical r.m.s. of the flux in a line-free channel or for 50 000 iterations, whichever condition was met first. The typical number of iterations was $\sim$ 20\,000 -- 30\,000. After the deconvolution process, each of the clean components was convolved with a Gaussian approximation of the dirty beam. The beam's full width at half maximum was 17$''\times 18.2''$ and $10.2''\times 10.2''$ for the naturally-weighted (NA) and robust-weighted (RW) data cubes respectively while the channel width was set to $dV=3.49$~km~s$^{-1}$\ for both. No Hanning smoothing was applied to either data cube. For each cube, the noise in a line-free channel is Gaussian distributed with a mean of $\mu\sim-0.02$~mJy~beam$^{-1}$ and a standard deviation of $\sigma\sim$~0.60~mJy~beam$^{-1}$ for the NA cube, while $\mu\sim-0.004$~mJy~beam$^{-1}$ and $\sigma\sim 0.58$~mJy~beam$^{-1}$ for the RW cube. \subsection{H\,{\sc i} Moment maps}\label{moment_maps} All H\,{\sc i} moment maps were generated using the Groningen Image Processing System (GIPSY, \citealt{gipsy}) together with the NA H\,{\sc i}\ data cube. The entire H\,{\sc i}\ data cube was smoothed, using a Gaussian convolving function, to a resolution of 34$''\times 36.4''$. A flux cut-off was applied to the smoothed cube at 2.5$\sigma$. For each channel of the smoothed data cube, any flux that was clearly not spatially connected to, or associated with the galaxy emission was flagged and removed. The blotted smoothed cube was then applied as a mask to the original $17''\times 18.2''$ cube. The H\,{\sc i} total intensity and 2nd-order moment maps were constructed from the zeroth- and second-order moments of the data cube respectively. A map in which each pixel is equal to $\sigma\sqrt{N}$, where $N$ is the number of un-blanked channels for a given line profile that would contribute to the moment maps, was created. Such a construction was permitted since no Hanning smoothing was applied to the H\,{\sc i}\ data cube. Dividing each pixel in the total intensity map by $\sigma\sqrt{N}$ produced a dimensionless signal-to-noise (S/N) map. The average intensity of all the pixels in the total intensity map that had a corresponding signal-to-noise ratio in the range 2.75 $\leq$ S/N $\leq$ 3.25 was used as an intensity cut-off for the total intensity map. This removed, in a robust manner, pixels from the outer disk that had low S/N values. This new H\,{\sc i}\ total intensity map was then used as a mask on other H\,{\sc i} maps. The method of using Gauss-Hermite polynomials to parameterise line profiles of galaxies was introduced by \citet{GHpolynomials}. While Gaussians are reasonable approximations to many realistic line profiles, higher order Gauss-Hermite polynomials are able to better capture profile asymmetries and non-Gaussian deviations. The Hermite method has already been successfully implemented by various authors \citep{GHpolynomials,noordermeer2007,THINGS_deblok}. We fitted a third-order Gauss-Hermite polynomial to each NA H\,{\sc i}\ data cube line profile to generate a so-called Gauss-Hermite H\,{\sc i}\ velocity field. Three filters were used simultaneously when fitting the profiles: 1) profiles with fitted peak fluxes below 2.5$\sigma$ were excluded; 2) profiles with a fitted line width less than the channel width were excluded; and 3) fitted profile peaks had to be within the velocity range of the data cube. \section{HI properties}\label{data_products} \subsection{Channel maps}\label{channel_maps} A grey-scale representation of the $17''\times18.2''$ resolution channel maps is shown in Fig. \ref{channels}. $\sigma\sim 0.6$~mJy in all channels. H\,{\sc i}\ structures are visible at large and small scales. At large scales, the emission clearly exhibits the usual pattern of a rotating disk. The H\,{\sc i}\ disk of NGC 2915 exhibits clear spiral structure (c.f. H\,{\sc i}\ total intensity map, Fig. \ref{mom0}). On small scales, the gas in the central regions forms two dominant over-densities in nearly all of the channels, with fainter emission in-between them. \begin{figure*} \centering \includegraphics[width=1.8\columnwidth, angle=0]{17as_channel_maps_23_45V2.eps} \caption{Channel maps of NA H\,{\sc i}\ data cube. Every second channel is shown. The heliocentric radial velocity (km s$^{-1}$) of each channel is shown in the upper left corner. The half-power-beam-width is shown in the bottom left corner. Grey-scale range is from -0.6 mJy beam$^{-1}$ to 10.0 mJy beam$^{-1}$. The r.m.s. noise in a channel is $\sim$ 0.6 mJy beam$^{-1}$.} \label{channels} \end{figure*} \addtocounter{figure}{-1} \begin{figure*} \centering \includegraphics[width=1.8\columnwidth, angle=0]{17as_channel_maps_47_69V2.eps} \caption{Continued.} \end{figure*} \subsection{Global profile}\label{global_profile} A global H\,{\sc i}\ profile, extracted from the blotted NA H\,{\sc i}\ data cube by summing the H\,{\sc i}\ emission in each velocity channel, is presented in Fig.~\ref{global_HI_profiles}. Using a synthesis array to estimate the total H\,{\sc i}\ flux of a galaxy can result in an underestimate due to missing short baselines. The shortest baseline used for our observations was 31~m (as part of the EW352 antenna configuration), implying that the observations are less sensitive to structures larger than $\sim$ 23$'$. \citet{meurer2}, however, estimated the H\,{\sc i} extent of NGC~2915 to be $\sim$ 19$'$. The amount of H\,{\sc i} missed in the global H\,{\sc i} profile is therefore expected to be small. In Fig.~\ref{global_HI_profiles} we compare our determination of the global H\,{\sc i}\ profile for NGC~2915 to the HIPASS global H\,{\sc i}\ profile from \citet{HIPASS_1000}. The HIPASS data are calibrated with 8$'$ spatial binning and 13 km s$^{-1}$ channel width. Due to the large HIPASS beam, as well as the fact that the Parkes telescope is a single dish telescope, the amount of flux missed in the HIPASS observations should be negligible. It is clear from Fig.~\ref{global_HI_profiles} that no significant amount of flux has been missed by our observations. The resulting global H\,{\sc i}\ profile is double-horned and asymmetric. From this profile we derived the profile widths, the systemic velocity, and the total H\,{\sc i} mass. The profile widths were calculated as the differences between the high and low velocities of the galaxy at 20$\%$ and 50$\%$ of the peak flux density (V$_{high}^{20\%}$, V$_{low}^{20\%}$, V$_{high}^{50\%}$, V$_{low}^{50\%}$ respectively). The profile widths were not corrected for inclination. Two systemic velocity estimates were determined: 1) from the global profile, at the midpoint of the two velocities at the 50$\%$ level, 2) from the equation \begin{equation} V_{sys}=0.25\times(V_{high}^{20\%}+ V_{low}^{20\%}+ V_{high}^{50\%}+ V_{low}^{50\%}). \label{vsysest} \end{equation} Finally, the total H\,{\sc i} mass was calculated as \begin{equation} M_{HI}=2.36\times 10^5\times D^2\times\int F dV, \end{equation} where $D$ is the distance to the galaxy in Mpc and $\int F dV$ is the total H\,{\sc i} line flux in Jy~km~s$^{-1}$. This determination of the total H\,{\sc i} mass assumes that the H\,{\sc i} is optically thin. Throughout this paper we adopt the $D\sim 3.78$ Mpc distance determination of \citet{karachentsev_catalog}. Table~\ref{HI_global_profile_table} lists our determinations of the above mentioned quantities. Assuming a distance of $D=5.1$~Mpc, \citet{meurer2} measured a total H\,{\sc i}\ mass of $M_{HI}\sim1.27\times 10^9$~$M_{\odot}$\ for NGC~2915. Using a distance of $D=3.78$~Mpc, their total mass estimate reduces to $M_{HI}\sim 7 \times10^8$~$M_{\odot}$, consistent with our estimate of $M_{HI}\sim 4.4\times 10^8$~$M_{\odot}$. \begin{table} \begin{center} \caption{Quantities derived from the global H\,{\sc i} profile.} \scriptsize{ \begin{tabular}{ccccc} \hline \hline \\ 1 & 2 & 3 & 4 & 5 \\ $M_{HI}$ & $W_{50}$ & $W_{20}$ & $V_{sys}^{50}$ & $V_{sys}^{\mathrm{Eqn. 1}}$ \\ (10$^8$ M$_{\odot}$) & (km s$^{-1}$) & (km s$^{-1}$) & (km s$^{-1}$) & (km s$^{-1}$) \\ \\ \hline \\ 4.4 & 186.8 & 170.0 & 460.5 & 460.8\\ \\ \hline \end{tabular}} \label{HI_global_profile_table} \\ \end{center} \textbf{Comments on columns:} Column 1: total H\,{\sc i} mass; Column 2: velocity width at 50$\%$ of the peak flux; Column 3: velocity width at 20$\%$ of the peak flux; Column 4: systemic velocity from W$_{50}$ midpoint; Column 5: systemic velocity from Eqn. \ref{vsysest}. \end{table} \begin{figure} \begin{centering} \includegraphics[angle=0,width=1.0\columnwidth]{globalHI_paperV2.eps} \caption{Global H\,{\sc i} profile from our observations (solid curve) and HIPASS (dashed curve). The dotted vertical line shows the systemic velocity as calculated using Eqn \ref{vsysest}.} \label{global_HI_profiles} \end{centering} \end{figure} \subsection{Total intensity map}\label{total_intensity} The full H\,{\sc i}\ total intensity map extracted from the NA H\,{\sc i}\ data cube is shown in the upper panel of Fig. \ref{mom0}. The edge of the H\,{\sc i}\ disk (the outer-most contour level) lies at $R\sim$~510$''$ and therefore extends out to $R\sim$~20 $B$-band scale lengths. For comparative purposes the IRAC 3.6~\micron\ image of the stellar disk is shown in Fig. \ref{IRAC_image}. To determine the photometric centre we fitted ellipses to three 3.6~\micron\ flux density annuli near the edge of the old stellar disk. A 3.6~\micron\ flux density contour at a level of 1.2~MJy~ster$^{-1}$, the average flux density of the second annulus, is shown in Fig.~\ref{IRAC_image}. The average flux densities of the other two annuli are 0.8 and 1.7~MJy~ster$^{-1}$. The average centre position of the ellipses fitted to the flux density annuli, $\alpha_{2000}$~=~09$^\mathrm{h}$~26$^\mathrm{m}$~12.611$^\mathrm{s}$, $\delta_{2000}$~=~$-76$$^{\circ}$~37$'$~37.80$''$, was used to estimate the position of the photometric centre (shown as a cross in Fig.~\ref{IRAC_image}). For comparison, the 2MASS determination of the centre of the system is $\alpha_{2000}$~=~09$^\mathrm{h}$~26$^\mathrm{m}$~11.53$^\mathrm{s}$, $\delta_{2000}$~=~$-76$$^{\circ}$~37$'$~34.80$''$ \citep{skrutskie_2006}. To facilitate length-scale comparisons between IR and H\,{\sc i}\ images, an ellipse with semi-major axis length $a=50''$, roughly equal to the $R$-band $R_{25}$ radius determined by \citet{meurer1}, is included in the figures. In their study of the H\,{\sc i}\ content of 73 late-type dwarf galaxies, \citet{WHISP_swaters} found that the ratio of the H\,{\sc i}\ extent to the optical diameter, defined as 6.4 disk scale lengths, is 1.8 $\pm$ 0.8 on average. The same ratio for NGC~2915 is $\sim 3.4$, close to double that value. \begin{figure} \centering \includegraphics[angle=0,width=1.0\columnwidth]{mom0_paperV2.eps} \caption{Grey-scale plot of the H\,{\sc i}\ total-intensity map (top panel) with a zoom-in of the central region (bottom panel). In both panels the intensity scale runs from 5~mJy~beam$^{-1}$~(white) to 0.2~Jy~beam$^{-1}$ (black). Contour levels run from 10~--~60~mJy~beam$^{-1}$ in steps of 10~mJy~beam$^{-1}$ and from 80~--~230~mJy~beam$^{-1}$ in steps of 20~mJy~beam$^{-1}$. The position of the photometric centre estimated using the 3.6~\micron\ image is marked with a cross. The hatched circles in the lower corners represent the half power beam width of the synthesised beam. The black and white ellipses have a semi-major axis length $a=50''$, an inclination $i=55^{\circ}$, a position angle $PA=317^{\circ}$ and a centre position equal to our determination of the photometric centre. The solid black lines represent the position-velocity slices shown in Fig.~\ref{pv_slices}.} \label{mom0} \end{figure} \begin{figure} \centering \includegraphics[angle=0,width=1.0\columnwidth]{irac_paperV2.eps} \caption{Grey scale plot of the 3.6~\micron\ IRAC Spitzer image. The intensity scale runs from -0.2~MJy~ster$^{-1}$~(white) to 2.0~MJy~ster$^{-1}$~(black). The single white contour is at a level of 1.2~MJy~ster$^{-1}$ and represents the average flux of one of three flux annuli used to estimate the position of the photometric centre. The estimated photometric centre is marked with a cross. The black ellipse is the same as that shown in Fig.~\ref{mom0}} \label{IRAC_image} \end{figure} Our high-resolution H\,{\sc i}\ observations resolve the central regions of NGC~2915 into two distinct H\,{\sc i}\ concentrations (Fig.~\ref{mom0}, lower panel). Their centres are separated by $\sim$~60$''$ (1.1~kpc) and have a combined mass of $\sim 1.9\times 10^7$~$M_{\odot}$. Also visible is a plume-like H\,{\sc i}\ feature to the North-West of these central H\,{\sc i}\ concentrations with a mass of $\sim 2.8\times 10^7$~$M_{\odot}$. The observed spiral structure of the outer disk is asymmetric: the arm beginning at the North-Western central H\,{\sc i}\ concentration can be traced over $\sim$ 200$^{\circ}$ in azimuthal angle while the arm beginning at the South-Eastern H\,{\sc i}\ concentration can be traced for only $\sim$ 120$^{\circ}$. From the H\,{\sc i} total intensity map an azimuthally averaged, inclination-corrected H\,{\sc i} surface density profile was constructed. Each ring has a width of $dR=17''$, a position angle (measured anti-clockwise from North to the receding major axis) of $PA=285^{\circ}$ and an inclination of $i=55$$^{\circ}$. The profile is shown in Fig. \ref{surf_dens_prof}. In an attempt to better constrain the H\,{\sc i}\ distribution of NGC~2915, an H\,{\sc i}\ total intensity map was extracted from the RW H\,{\sc i}\ data cube. For the sake of brevity this map is not shown here, yet it suffices to say that the higher spatial resolution confirms the above-mentioned nature of the central H\,{\sc i}\ concentrations. However, the very low S/N ratio in the outer parts of this map yields it unsuitable for the study of low column density H\,{\sc i}\ features, with only the central-most H\,{\sc i}\ features standing out about the noise. This map does not allow for a more accurate study of the H\,{\sc i}\ distribution and is therefore not discussed further. \begin{figure} \centering \includegraphics[angle=0,width=1.0\columnwidth]{HI_surface_density_profileV2.eps} \caption{H\,{\sc i}\ surface density radial profile. All surface densities have been inclination-corrected. Error bars represent the r.m.s. spread of the surface densities in each azimuthally averaged ring.} \label{surf_dens_prof} \end{figure} \subsection{Velocity field}\label{vel_field} The third-order Gauss-Hermite velocity field extracted from the NA H\,{\sc i}\ data cube is shown in Fig.~\ref{mom1}. The blank pixels in this map are those that were rejected by the fitting routine according to the criteria mentioned in Sec. \ref{moment_maps}. \citet{THINGS_deblok} showed that the Gauss-Hermite parameterisation of a line profile gives a robust estimate of the peak-associated velocity of the profile. Inner contours running approximately parallel to the kinematic minor axis are caused by $V(R)\propto R$ rotation while the outer contours running radially away from the centre are due to a constant rotation speed. The overall ``S-shaped'' distortion of the contours is indicative of a kinematic warp within the disk. Small wiggles along the outer contours are caused by streaming gas motions. The sharp kinks at inner radii suggest the presence of non-circular velocity components within the gas. Finally, differences in the shapes of the iso-velocity contours on the receding and approaching halves of the galaxy suggest a certain degree of kinematic lopsidedness within the galaxy. A third-order Gauss-Hermite velocity field was also extracted from the RW H\,{\sc i}\ data cube. Due to the high noise levels in the RW H\,{\sc i}\ cube, most line profiles were rejected by the fitting filters and hence not fitted. The general characteristics of the fitted peaks of the remaining profiles are very similar to those of the NA velocity field, differing on average by $\sim 5$~km~s$^{-1}$. Rather than showing this H\,{\sc i}\ velocity field here, a tilted ring model is fitted to it in Sec.~\ref{TR_output} to demonstrate that the resulting rotation curve is very similar to the one derived using the NA third-order Gauss-Hermite velocity field. Motivated by this result is the decision to use only the NA velocity field for our kinematic analyses. \begin{figure} \centering \includegraphics[angle=0,width=1.0\columnwidth]{mom1_paperV2.eps} \caption{Top panel: Grey-scale plot of the third-order Gauss-Hermite H\,{\sc i} velocity field. Bottom panel: A zoom-in of the central region. The intensity scale runs from 300 -- 600 km~s$^{-1}$. Contours are separated by 15 km s$^{-1}$ with the thick contour marking the systemic velocity at 465 km s$^{-1}$ derived by fitting a tilted ring model to the velocity field. White and black contours represent receding and approaching halves of the galaxy respectively. The hatched circles in the lower corners represent the half power beam width of the synthesised beam. The ellipse is the same as that shown in Fig. \ref{IRAC_image}.} \label{mom1} \end{figure} \subsection{Second-order moment map}\label{disp_map} The second-order H\,{\sc i}\ moment map extracted from the NA H\,{\sc i}\ data cube is shown in Fig.~\ref{mom2}. The highest second-order moments are found near the stellar core of the galaxy. The radial profile of the second-order moment map is shown in Fig.~\ref{disp_prof}. If the second-order moments are interpreted as a measure of the gas velocity dispersion, then the average outer disk dispersion of $\sigma_{gas}\sim$ 10 km s$^{-1}$ is close to the constant gas velocity dispersion value of $\sigma_{gas}=11$~km~s$^{-1}$\ found by \citet{leroy_THINGS} for a sample of THINGS dwarf galaxies. \begin{figure} \centering \includegraphics[angle=0,width=1.0\columnwidth]{mom2_paperV2.eps} \caption{Grey-scale plot of the second-order H\,{\sc i} moment map. The grey scale runs from 6~km~s$^{-1}$~(white) to 30~km~s$^{-1}$~(black). Contour levels are at 6, 9, 12, 18, 22, 26, 28, 30~km~s$^{-1}$. The hatched circle in the lower left corner represents the half power beam width of the synthesised beam. The ellipse is the same as that shown in Fig. \ref{IRAC_image}.} \label{mom2} \end{figure} \begin{figure} \begin{centering} \includegraphics[width=1.0\columnwidth]{HI_disp_profile_paperV2.eps} \caption{Second-order H\,{\sc i} moment map radial profile. Error bars represent the r.m.s. spread of the 2nd-order moments in each azimuthally averaged ring.} \label{disp_prof} \end{centering} \end{figure} \subsection{Non-Gaussian line profiles}\label{line_profiles} Position-velocity slices taken though the centre of the NA H\,{\sc i}\ data cube betray the presence of broad and multi-component line profiles at inner and outer radii. Figure \ref{pv_slices} shows two position-velocity slices, the positions of which are shown in the H\,{\sc i}\ total intensity map (Fig. \ref{mom0}, top panel). The first slice was taken at a position angle of 322$^{\circ}$, through both of the central H\,{\sc i}\ concentrations. This slice suggests that the galaxy contains a fast-rotating gas component near its centre, clearly manifesting itself as a sharp velocity spike with a peak velocity of $V\sim$ 550 km~s$^{-1}$. Line profiles near the centre of this position-velocity slice show significant departures from Gaussianity. This position-velocity slice also reveals asymmetric line profiles on the receding side of the galaxy, seen in the upper-left quadrant ($V>V_{sys}=465$~km~s$^{-1}$, angular offset $<0$) as an ``H\,{\sc i}\ beard'' that is lagging in velocity relative to the main disk. The beard emission spatially coincides with the plume-like H\,{\sc i}\ feature seen in the H\,{\sc i}\ total intensity map. The estimated mass of this morphological feature, $\sim 2.8\times 10^7$~$M_{\odot}$, is 5.6$\%$ of total H\,{\sc i}\ mass. One interpretation of this kinematically anomalous gas component is that it is associated with pristine gas being accreted from the nearby inter-galactic space onto the outer disk of NGC 2915. If confirmed, this would be the first evidence of such an event in a dwarf system like NGC~2915. The second position-velocity slice was extracted almost parallel to the minor axis, through the South Eastern central H\,{\sc i}\ concentration which contains many double--component profiles. Line profiles in this central region are split by $\sim~30$~km~s$^{-1}$\ on average, implying that the sharp rise seen toward the centre of the second-order H\,{\sc i}\ moment map (Fig.~\ref{mom2}) over-estimates the true H\,{\sc i}\ velocity dispersions of the inner disk. Since the disturbed H\,{\sc i}\ line profiles are located within $R\sim 150''$ of the centre of NGC 2915, it is plausible that the central gas dynamics of NGC~2915 are largely dictated by the stellar winds of its young stellar population. In a forthcoming paper (Elson et al. 2010, in prep.) we present detailed modeling results of the central H\,{\sc i}\ dynamics. \begin{figure} \begin{centering} \includegraphics[angle=0,width=1.0\columnwidth]{non_gauss_lines_paperV2.eps} \caption{Position-velocity diagrams extracted along 1) a position angle of 322$^{\circ}$ (top) and 2) a position angle of 220$^{\circ}$ (bottom). The slice positions are shown in the H\,{\sc i}\ total intensity map (Fig.~\ref{mom0}). The thickness of the slices is that of a single pixel in the data cube (2.5$''$) and the velocity resolution is 3.5 km s$^{-1}$. The intensity scale runs from --3$\sigma$ to (white) 40$\sigma$ (black). Contours start at 2$\sigma$ in steps of 3$\sigma$. The dashed horizontal line marks the systemic velocity at $V_{sys}=465$~km~s$^{-1}$.} \label{pv_slices} \end{centering} \end{figure} \section{H\,{\sc i}\ dynamics}\label{vrot} The H\,{\sc i} line data allow us to study the H\,{\sc i} gas dynamics. In this section we derive a rotation curve for NGC~2915. \subsection{Tilted ring model} The standard method of producing a rotation curve involves fitting a tilted ring model to a velocity field \citep{begemann_phd_thesis}. NGC~2915's intermediate inclination of $i\sim$55$^{\circ}$ makes it a well-suited candidate for such an analysis. The method involves modeling the galaxy as a set of concentric rings within which the gas orbits about the kinematic centre. Each ring has a set of defining parameters: central coordinates $X_c$ and $Y_c$, inclination $i$, systemic velocity $V_{sys}$, position angle $PA$ (in the sky plane\footnote{Measured anti-clockwise from North to the receding semi-major axis.}), and the rotation velocity $V_{rot}$ of the material within the ring. In the case that expansion velocities within the disk are ignored, the standard algorithm carries out a least squares fit to \begin{equation} V_{los}(x,y)=V_{sys}+V_{rot}\sin(i)\cos(\theta), \label{vlos} \end{equation} where $V_{los}$ is the line-of-sight velocity, $x$ and $y$ are rectangular coordinates on the sky and $\theta$ is the angle from the major axis in the galaxy plane. $\theta$ is related to the position angle $PA$ in the sky plane by \begin{eqnarray} \cos(\theta)&=&{-(x-X_c)\sin(PA)+(y-Y_c)\cos(PA)\over r},\\ \sin(\theta)&=&{-(x-X_c)\cos(PA)+(y-Y_c)\sin(PA)\over r\cos(i)}, \label{vlos} \end{eqnarray} where $r$ is the radius of the ring in the galaxy plane. The GIPSY task ROTCUR \citep{begemann_phd_thesis} was used to fit tilted ring models to the third-order Gauss-Hermite H\,{\sc i} velocity field extracted from the NA H\,{\sc i}\ data cube. Rings width of $dr=17''$ were used to ensure that adjacent rings were largely independent of one another. When fitting Eqn.~\ref{vlos} to the data, each datum was weighted by $|\cos(\theta)|$ so that points closer to the major axes held more weight. All points within 10$^{\circ}$ of either side of the minor axes were excluded from the fit. Both sides of the galaxy were used for the fitting. \subsection{Fitting procedure} Each tilted ring should be centred on the dynamical centre of the galaxy, assuming it is constant, which is usually estimated by the least-squares fitting algorithm together with the other tilted ring parameters. In the case of NGC~2915, however, the kinematic centre determinations from ROTCUR varied with radius in a non-systematic manner. We therefore used the photometric centre of the 3.6~\micron\ emission as the kinematic centre. Using the photometric and kinematic centres interchangeably in this manner is reasonable if one assumes that the stars lie at the bottom of the gravitational potential. We carried out a single ROTCUR iteration with all parameters allowed to vary freely. The resulting $X_c$ and $Y_c$ values are shown as filled black circles in panels A and B respectively of Fig. \ref{TR_results}. These fitted $X_c$ and $Y_c$ positions deviate on average by 8.3$''$~$\pm$~$0.3''$ and 7.2 $''$~$\pm$ 0.25$''$ (152~pc and $\sim$ 132~pc) respectively from the photometric centre, thereby placing them well within a single half power beam width. This average deviation is consistent with the results of \citet{trachternach_THINGS} who found that approximately 50$\%$ of their kinematic centres derived for $\sim$~1000 individual tilted rings fitted to the THINGS H\,{\sc i}\ velocity fields differed by less than a beam width from their best centre estimates. In the case of the THINGS sample, one beam width corresponds to $\sim$~200~pc on average. Throughout the tilted ring fitting procedure, $X_c$ and $Y_c$ were therefore each fixed to the position of the photometric centre (solid lines in panels A and B of Fig. \ref{TR_results}) \begin{figure*} \begin{centering} \includegraphics[width=2.0\columnwidth]{TR_resultsV2.eps} \caption{Radial variations of tilted ring parameters fitted to the third-order Gauss-Hermite velocity field. Panels are A: central X positions, B: central Y positions, C: position angles, D: inclinations, E: systemic velocities, and F: circular rotation velocities. Filled circles show the radial variations just before the parameter profiles were smoothed and fixed in order to derive the final rotation curves. The smoothed profiles used to construct models SI and CI are shown as solid curves except for the model CI inclination profile which appears as a dashed curve (offset by -2$^{\circ}$ for the sake of clarity). Panel F shows the final rotation curves derived for models SI and CI (open and filled squares respectively). For the sake of clarity the rotation curves have been offset by 8.5$''$. Error bars represent the r.m.s. spread of the velocities in a given ring.} \label{TR_results} \end{centering} \end{figure*} The remaining tilted ring parameters ($V_{sys}$, $PA$, $i$, $V_{rot}$) were derived iteratively. We first allowed all parameters to vary freely. This gave a general feel for the parameter behaviour with $r$. Little scatter was seen in the radial run of $V_{sys}$, which was then fixed to the average value of 465~km~s$^{-1}$\ in subsequent iterations (compare with the estimate of $V_{sys}=460.8$~km~s$^{-1}$\ derived from the global H\,{\sc i}\ profile using Eqn.~\ref{vsysest}). The $PA$, $i$, and V$_{rot}$ parameters were allowed to vary freely in the subsequent iterations. The $PA$ and $i$ runs showed relatively little scatter except for a clear change from outer to inner disk beginning at $R$ $\sim$ 240$''$ where the inclination value changed from $i\sim$ ~55$^{\circ}$ to $i\sim$~75$^{\circ}$. Our kinematically derived position angles were consistent with the position angles of ellipses fitted to flux contours in the H\,{\sc i} total intensity map. Several more iterations with various combinations of fixed and free parameters were carried out in order to check the stability of the radial parameter profiles. The sharp rise in $i$ for R $\lesssim$ 240$''$ always remained with little scatter. \begin{figure} \begin{centering} \includegraphics[width=1\columnwidth]{vrot_app_recV2.eps} \caption{Rotation curves from tilted ring model SI derived separately for the approaching (filled circles) and receding (open circles) sides. For the sake of clarity the rotation curve of the approaching side has been offset by 8.5$''$. Error bars represent the r.m.s. spread of the velocities in a given ring.} \label{2vrots} \end{centering} \end{figure} \begin{figure*} \begin{centering} \includegraphics[width=1.85\columnwidth]{pv_overlays_SI_paperV2.eps} \caption{Position-velocity slices though the kinematic centre of NGC~2915 at various position angles. Overlaid on each slice is the rotation curve of tilted ring model SI projected using the smoothed parameter radial profiles as shown in Fig. \ref{TR_results}. The thickness of the position-velocity slices is that of a single pixel in the data cube (2.5$''$) and the velocity resolution is 3.5 km s$^{-1}$. The dashed horizontal line shows the systemic velocity at 465 km s$^{-1}$ adopted for our kinematic analyses. Contours start at 2$\sigma$ in steps of 2.5$\sigma$.} \label{pv_overlays} \end{centering} \end{figure*} For the final iteration, the $X_C$, $Y_C$ and $V_{sys}$ parameters were fixed to values that were constant with $r$ while the inclination and position angle radial profiles were smoothed and then fixed. The smoothed versions of the profiles are shown as solid curves in panel A, B, C, D, E in Fig. \ref{TR_results} while the filled circles represent the parameter values just before they were smoothed and fixed. Smoothing the $PA$ and $i$ radial distributions did not affect the final rotation curve in any noticeable manner. As \citet{THINGS_deblok} point out, the smoothing only affects the point-to-point scatter of the profiles and in no way affects the resolution of the radial distributions. To derive the final rotation curve, only $V_{rot}$ was allowed to vary. One of the main sources of uncertainty in the derivation of the inner rotation curve is the rise in inclination of $\sim$~20$^{\circ}$\ towards inner radii preferred by the tilted ring models. It is not clear whether this rise is due to a genuine kinematic warp, or whether it is merely an artefact of the non-Gaussian line profiles near the galaxy's centre. We therefore derived two final tilted ring models, one with and one without the rise in inclination (solid and dashed curves in panel D of Fig. \ref{TR_results} respectively). Both models used the same $X_c$, Y$_c$, $V_{sys}$ and $PA$ radial profiles. We hereafter refer to these two tilted ring models as models SI (Steep Inclination) and CI (Constant Inclination). In panel F of Fig. \ref{TR_results}, the rotation curves of both models are presented. In each of the panels A~-~E in Fig. \ref{TR_results}, the error bars represent the formal least squares errors from the ROTCUR fitting routine. In panel F, the error bars represent the r.m.s. spread of the velocities in a single ring. \subsection{Rotation curve}\label{TR_output} The final tilted ring models fitted to the third-order Gauss-Hermite velocity field are presented in Fig. \ref{TR_results}. While the stellar disk seen in the 3.6~\micron\ image (Fig.~\ref{IRAC_image}) extends as far as $R\sim 100''$, the inner portion of the rotation curve rises as $V(R)\propto R$ out to $R\sim 150''$. Beyond this radius the rotation velocity remains almost constant. For model SI, the average rotation velocity for $R\ge 187''$ is $V=81.9$~$\pm$~1.6~km~s$^{-1}$. Despite using different inclination profiles to derive the two final rotation curves, the maximum absolute difference between them is $V=9.3\pm12.2$~km~s$^{-1}$\ at $R=136''$. The average absolute difference for $R\le 187''$ is $V=4.3\pm 4.8$km~s$^{-1}$. The resulting two rotation curves are thus very similar to one another. For each tilted ring model we constructed separate rotation curves for the approaching and receding sides of the galaxy. In the case of model SI (Fig.~\ref{2vrots}), the rotation curve differs significantly between the two sides of the galaxy. Within the steeply rising portion, differences of $\sim$~20~km~s$^{-1}$\ are observed as well as at $R\sim 290''$. The same is true in the case of model~CI. \citet{swaters_lopsided} state that galaxies with asymmetric global profiles often have rotation curves that are more slowly rising on one side of the galaxy than the other. This statement seems true in the case of NGC~2915. To check that the two tilted ring models are consistent with the data, we used the smoothed radial parameter profiles to project the rotation curves onto various position-velocity slices extracted from the H\,{\sc i}\ data cube. The overlays for tilted ring model SI are presented in Fig~\ref{pv_overlays}. The projected rotation curves of the two tilted ring models are almost identical to one another with model SI fitting well the high-intensity regions of the position-velocity slices. Given a set of fixed radial profiles, the ROTCUR routine will adjust the rotation curve so that the \emph{line-of-sight} velocities best match the data. Thus although the circular rotation curves (i.e. panel F of Fig.~\ref{TR_results}) differ, the good agreement between the projected rotation curves is not surprising. Using the smoothed, fixed radial parameter profiles of model SI and ring widths of $dr=10''$, a tilted ring model was also fitted to the third-order Gauss-Hermite velocity field extracted from the RW H\,{\sc i}\ cube. The resulting rotation curve is compared to the rotation curve of model SI in Fig.~\ref{10as_17as_vrots}. The figure demonstrates that for $R\lesssim 250''$, the derived rotation curves are very similar to one another. Beyond $R\sim 250''$ the filling factor of the RW H\,{\sc i}\ velocity field becomes too low for meaningful derivations of the rotation velocity. For these reasons, only the rotation curves of tilted ring models CI and SI are used as mass model inputs in Sec.~\ref{mass_modeling}. \begin{figure} \begin{centering} \includegraphics[width=1\columnwidth]{10as_17as_vrotsV2.eps} \caption{The model SI rotation curve (open circles) compared to the rotation curve derived by fitting a tilted ring model to the H\,{\sc i}\ velocity field extracted from the RW H\,{\sc i}\ data cubes (black filled circles). Error bars represent the r.m.s. spread of the velocities in a single ring of the tilited ring model fitted to the RW velocity field. For the sake of clarity, error bars are not shown for the model SI rotation curve.} \label{10as_17as_vrots} \end{centering} \end{figure} \section{Mass modeling}\label{mass_modeling} Dwarf galaxies such as NGC 2915 serve as useful probes of dark matter (DM). In this section we derive two mass models for NGC 2915, one for each of the tilted ring models CI and SI. The gravitational potential within a galaxy is determined by the combined gravitational potentials of all the mass components. The derived 3.6~\micron\ surface brightness profile of NGC 2915 shows little deviation from an exponential light distribution at any radius. We therefore concluded that the galaxy does not contain a significant central stellar bulge. The total mass was treated as the sum of the masses of the stellar and gas disks as well as the DM halo. Since $M\propto V^2$ at a given $R$, the rotation curves of the individual mass components, when summed in quadrature, will yield the square of total rotation curve, $V_{tot}$: \begin{equation} V_{tot}^2=\alpha_{gas} V_{gas}^2+\Upsilon_* V_*^2+V_{DM}^2, \label{vtot} \end{equation} where $V_{gas}$, $V_*$, and $V_{DM}$ are the contributions to the total rotation curve of the gas, the stars and the DM respectively. $V_{tot}$ should match the observed rotation curve. $\Upsilon_*$ is the stellar mass-to-light ratio used to convert the observed distribution of star light to a stellar mass distribution while $\alpha_{gas}=1.37$ is the scaling factor used for all mass models to take into account the contribution of helium to the rotation curve. \subsection{Gas and stellar distributions}\label{bary_vrots} We used the GIPSY task ROTMOD \citep{rotmod} to convert the observed mass distributions into rotation curves. Each of the fitted tilted ring models (Sec. \ref{TR_output}) was used together with the H\,{\sc i}\ total intensity map to yield an H\,{\sc i}\ surface density profile. These profiles were used as ROTMOD input which, assuming an infinitely thin gas disk, yielded two different H\,{\sc i}\ rotation curves. Ellipses fitted to 3.6~\micron\ surface brightness isophotes suggest a constant inclination of $i\sim 55^{\circ}$ for the old stellar population. Using this inclination and a position angle of 306$^{\circ}$, a single surface density profile for this mass component was derived. This surface density profile was extrapolated out to $R=510''$ to match the radial extent of the observed rotation curve, and then converted into a stellar rotation curve assuming a $\sec$h$^2$ vertical distribution \citep{van_der_kruit_1981} and $h/z_0=5$, yielding $z_0=0.12$~kpc. In principle the DM density distribution of a suitably rotating system can be determined knowing the contributions to the observed rotation curve from the gas and the stars. The H\,{\sc i}\ rotation curve is directly related to the observed H\,{\sc i}\ distribution, assuming that the H\,{\sc i}\ is optically thin. The situation is not as simple for the stellar component of a galaxy, however, since a stellar mass-to-light ratio is needed to link the stellar light and mass distributions. Determining the stellar mass-to-light ratio from the rotation curve alone is problematic. Given an observed rotation curve, equally well-fitted mass models can be obtained for a range of stellar mass-to-light ratios \citep[e.g.][]{van_albada_1985,swaters_thesis}. This degeneracy therefore leads to uncertainties in stellar mass-to-light ratios determined solely by rotation curves. \citet{bell_dejong} used spectroscopic and photometric spiral galaxy evolution models to show that a reasonable correlation exists between the optical and infrared stellar mass-to-light ratios and the colours of integrated stellar populations. They further showed that the relative trends between model stellar mass-to-light ratios and colours are fairly insensitive to uncertainties in stellar population and galaxy evolution modeling. \citet{THINGS_oh} used stellar population synthesis models (together with a scaled Salpeter 1955 IMF) to determine a relationship between the stellar mass-to-light ratios in the 3.6~\micron\ and $K$ bands ($\Upsilon_*^{3.6}$ and $\Upsilon_*^K$ respectively): \begin{equation} \Upsilon_*^{3.6}=0.92\times \Upsilon_*^K-0.05. \label{emp_ML} \end{equation} From \citet{bell_dejong}, the relation between the $B-V$ colours and the $K$-band mass-to-light ratio is \begin{equation} \log_{10}\left(\Upsilon_*^K\right)=b^K\times (B-V)+a^k, \label{k-colour} \end{equation} where $a^K$ and $b^k$ for varying colours and metallicities are provided in their Table 4. By combining these two equations we can estimate the stellar mass-to-light ratio at 3.6~\micron. We use $a^K=-0.59$ and $b^K=0.60$ for $B-V$ colours as used by \citet{THINGS_oh} for IC 2574, the justification being that IC~2574 is also a low metallicity, star-forming dwarf galaxy with $B-V\sim 0.6$, much like NGC 2915. $B-V\sim 0.3 \pm 0.2$ for NGC~2915 \citep{RC3,schlegel} thereby yielding (from Eqn.~\ref{k-colour}) $\Upsilon_*^K\approx 0.39$. Substituting this value into Eqn. \ref{emp_ML} yields $\Upsilon_*^{3.6}\approx 0.3$. \citet{leroy_THINGS}, applying the \citet{bell_et_al_2003} relation between $B-V$ colour and $\Upsilon_*^K$ and assuming a \cite{kroupa_IMF} IMF, found $\Upsilon_*^K=~0.48-0.60$ for the $K-$band stellar mass-to-light ratios of all of the THINGS galaxies in their sample. \subsection{Dark matter halos} The way in which dark matter is distributed on galactic length-scales has been a topic of much debate. Over the past decade, the constant increase in computational power has allowed detailed $N$-body simulations of the clustering properties of cold dark matter particles to be carried out. One of the first large-scale studies was carried out by \citet{NFW} in which they studied the equilibrium density profiles of dark matter halos in a hierarchial galaxy clustering scenario. Their results suggested the existence of a ``universal dark matter density profile'' \citep{NFW_1997} which is independent of clustering scale, mass or size as well as the power spectrum of initial fluctuations. Their parameterisation of this universal profile is \begin{equation} {\rho(r)_{NFW}\over \rho_{crit}}= {\delta_c \over {(r/r_s)(1+r/r_s)^2}} \end{equation} where $\delta_c$ is a measure of the density of the Universe at the time of collapse of the DM halo, $r_s$ is the characteristic scale radius and $\rho_{crit}=3H^2/8\pi G$ is the critical density required for closure. This profile scales as $r^{\alpha}$, with $\alpha=~-3$ and $\alpha=~-1$ at large and small radii respectively, thereby predicting extremely steep inner density profiles of DM halos, known as central cusps. The Aquarius Project \citep{aquarius_subhalos} recently completed a set of high-resolution numerical simulations to study the formation and structure of galaxy-sized DM halos in the standard $\Lambda$CDM cosmology. \citet{navarro_2008} reported that the spherically averaged density profile of the six different-sized simulated galaxy halos becomes progressively shallower inwards and, at the innermost resolved radius, has a logarithmic slope $\alpha = \mathrm{d}\ln \rho/\mathrm{d}\ln r \gtrsim -1$. These predictions of cuspy inner DM density profiles can be tested against real, dark-matter-dominated galaxies. If cuspy NFW halos are found, constraints can be put on halo concentrations and cosmological parameters. The rotation curve resulting from an NFW density profile is \begin{equation} \left({V_{NFW}\over V_{200}}\right)^2={\ln (1+cx)-cx/(1+cx)\over x\ln(1+cx)-c/(1+c)}, \end{equation} where $V_{200}$ is the circular rotation speed at $r_{200}$, the radius at which the density of the DM halo equals 200 times the critical density, $\rho_{crit}$; $x=~r/r_{200}$ is the radius in units of the virial radius and $c=r_{200}/r_s$ is the ``concentration'' parameter of the halo. Low surface brightness dwarf galaxies are thought to be dark-matter-dominated down to small radii \citep{deblok_mcgaugh_1996} and can therefore be more directly compared to the cold dark matter simulations. When this is done, it is often the case that the density profiles of simulated DM halos are too steep to fit the rotation curves of dwarf galaxies \citep{moore_1994,flores_primack,deblok_mcgaugh_1996,marchesini_2002}. The distribution of the DM in these dwarfs is instead consistent with approximately constant inner dark matter densities. A particular dark matter density profile that describes the DM distribution of dwarf systems well is that of the pseudo-isothermal sphere \citep{galactic_dynamics}: \begin{equation} \rho(r)_{ISO}=\rho_0\left(1+\left({r\over r_c}\right)^2\right)^{-1}, \end{equation} where $\rho_0$ is the central DM density and $r_c$ is the core radius. This parameterisation, which has no particular physical justification, has a constant density core. The rotation curve resulting from such a density profile is \begin{equation} V_{ISO}=\left(4\pi G \rho_0 r_c^2\left[1-{r_c\over r}\arctan\left(r\over r_c\right)\right]\right)^{1/2}. \end{equation} \subsection{Fitted models} We used the GIPSY task ROTMAS to construct mass models for NGC~2915. The task subtracts from the observed rotation curve the scaled rotation curves of the stellar and gas disks. It then fits to the residuals a rotation curve corresponding to a particular parameterisation of the DM halo. When matching the total and observed rotation curves, the points were uniformly weighted to ensure that each point in the rotation curve contributed equally to the fit. This essentially reduced the fitting procedure to a non-weighted least-squares fitting method. As previously mentioned, each of the rotation curves from the two tilted ring models fitted to the H\,{\sc i}\ velocity field were used as mass modeling input. For each of these mass models we carried out fits for a pseudo-isothermal sphere and an NFW halo. Upper and lower limits for the stellar mass-to-light ratio were determined by scaling the contribution of the stellar rotation curve to the total rotation curve. Under the so-called ``maximum disk assumption'', the stellar rotation curve is scaled up to contribute maximally to the observed rotation curve at inner radii. Such a scaling usually allows the stellar disk to explain most of the inner rotation curve \citep{van_albada_1985,swaters_thesis}. Under the ``minimum disk assumption'', the contribution of the stellar disk is made as low as possible (often zero) while still allowing a good fit of the total rotation curve to the observed rotation curve. Each of the halo fits was divided into 3 sub-cases: A fit 1) with $\Upsilon_*$ fixed to our predetermined value of 0.3; 2) using a maximum disk assumption; and 3) using a minimum disk assumption. For the minimum disk cases, $\Upsilon_*$ was fixed to zero (no contribution from the stellar disk to the observed rotation curve). For the maximum disk cases, regardless of the DM halo parameterisation, $\Upsilon_*$ was fixed to 0.64 and 0.44 for mass models SI and CI respectively. In all models, $\alpha_{gas}$ was fixed to 1.37. Thus, six mass models were fitted to each of the 2 rotation curves derived from the tilted ring modeling. \subsection{Results} \begin{table*} \begin{center} \caption{Mass modeling results} \label{mass_modeling_results} \scriptsize{ \begin{tabular}{cccccccc} \hline \hline \\ &&1 & 2 & 3 & 4 & 5 &6\\ \\ &DM halo & $ISO$ & $ISO_{max}$ & $ISO_{min}$ & $NFW$ & $NFW_{max}$ &$NFW_{min}$\\ \\ \hline \\ &\textbf{Model SI}\\ \\ 1&$\chi^2_{red}$ & 38.1 & 59.8 & 28.2 & 62.5 & 93.4 & 44.7 \\ 2&r.m.s. (km s$^{-1}$) & 0.56 & 1.30 & 0.79 & 0.75 & 1.37 & 0.40\\ 3& $\Upsilon_*$ ($M_{\odot}/L_{\odot}$)& 0.3 & 0.64 & 0.0 & 0.3 & 0.64 & 0.0\\ 4& $\rho_0$ (10$^{-3}$ $M_{\odot}$ pc$^{-3}$) & 100.6 $\pm$ 23.3 & 56.9 $\pm$ 17.8 & 178.7 $\pm$ 33.0 & ... & ... & ...\\ 5& $R_c$ (kpc) & 1.3 $\pm$ 0.2 & 1.7 $\pm$ 0.4 & 0.9 $\pm$ 0.1 & ... & ... & ...\\ 6& $V_{200}$ (km s$^{-1}$) & ... & ... & ... & 69.6 $\pm$ 7.9 & 83.5 $\pm$ 20.1 & 62.7 $\pm$ 4.1\\ 7& $C$ & ... & ... & ... & 11.7 $\pm$ 2.2 & 7.9 $\pm$ 2.7 & 15.6 $\pm$ 2.0\\ \\ &At last measured pt:\\ \\ 8& $M_{tot}$ (10$^{8}$ $M_{\odot}$) & 151.4 & 151.3 & 151.6 & 151.1 & 154.2 & 148.3\\ 9& $M_{HI}/M_{tot}$ (10$^{-3}$) & 43.84 & 43.85 & 43.79 & 43.92 & 43.03 & 44.76\\ 10& $M_{DM}/M_{tot}$ (10$^{-2}$) & 92.94 & 89.91 & 95.62 & 92.93 & 90.10 & 95.52\\ 11& $M_{DM}/M_{lum}$ & 6.68 & 4.47 & 21.84 & 6.67 & 4.57 & 21.34\\ 12& $M_{tot}/L_B$ ($M_{\odot}/L_{\odot}$) & 140.72 & 140.69 & 140.91 & 140.47 & 143.37 & 137.9\\ \\ \hline \\ & \textbf{Model CI}\\ \\ 1& $\chi^2_{red}$ & 46.7 & 53.6 & 37.5 & 78.1 & 89.1 & 61.2\\ 2& r.m.s. (km s$^{-1}$) & 0.74 & 1.00 & 0.27 & 1.03 & 1.28 & 0.66\\ 3& $\Upsilon_*$ ($M_{\odot}/L_{\odot}$)& 0.3 & 0.44 & 0.0 & 0.3 & 0.44 & 0.0\\ 4& $\rho_0$ (10$^{-3}$ $M_{\odot}$ pc$^{-3}$) & 127.9 $\pm$ 34.9 & 103.2 $\pm$ 31.1 & 227.6 $\pm$ 50.1 & ... & ... & ...\\ 5& $R_c$ (kpc) & 1.10 $\pm$ 0.2 & 1.2 $\pm$ 0.2 & 0.8 $\pm$ 0.1 & ... & ... & ...\\ 6& $V_{200}$ (km s$^{-1}$) & ... & ... & ... & 66.4 $\pm$ 8.3 & 70.3 $\pm$ 11.5 & 60.7 $\pm$ 4.7\\ 7& $C$ & ... & ... & ... & 12.9 $\pm$ 2.8 & 11.2 $\pm$ 2.9 & 17.0 $\pm$ 2.5\\ \\ & At last measured pt:\\ \\ 8& $M_{tot}$ (10$^{8}$ $M_{\odot}$) & 138.8 & 138.3 139.8 & 137.6 & 138.4 136.3\\ 9& $M_{HI}/M_{tot}$ (10$^{-3}$) & 47.29 & 47.43 & 46.92 & 47.66 & 47.42 & 48.16\\ 10& $M_{DM}/M_{tot}$ (10$^{-2}$) & 92.34 & 91.13 & 95.31 & 92.28 & 90.92 & 95.18\\ 11& $M_{DM}/M_{lum}$ & 6.11 & 5.15 & 20.31 & 6.06 & 5.01 & 19.77\\ 12& $M_{tot}/L_B$ ($M_{\odot}/L_{\odot}$) & 129.00 & 128.62 & 130.01 & 127.99 & 128.64 & 126.68\\ \end{tabular}} \end{center} \begin{flushleft} \textbf{Comments on rows:} Row~1:~Reduced $\chi^2$ goodness-of-fit statistic; Row~2:~r.m.s. of the difference between the observed and total rotation curves; Row~3:~stellar mass-to-light ratio; Row~4:~central density for pseudo-isothermal sphere; Row~5:~core radius for pseudo-isothermal sphere, Row~6:~circular rotation speed at virial radius for NFW halo; Row~7:~concentration parameter for NFW halo; Row~8:~dynamical mass; Row~9:~H\,{\sc i}\ to total mass ratio; Row~10:~Dark to total mass ratio; Row~11:~Dark to luminous mass ratio; Row~12:~Total mass to $B$-band light ratio.\\ \textbf{Comments on columns:} For each model, columns 1 and 4 show the mass modelling results for the case in which $\Upsilon_*$ is fixed to the predetermined value of 0.3. Columns 2 and 5 show the results for the maximum disk case while columns 3 and 6 are for the minimum disk case. \end{flushleft} \end{table*} \begin{figure*} \begin{centering} \includegraphics[angle=0,width=2\columnwidth]{rotmas_results_paperV2.eps} \caption{Pseudo-isothermal sphere (ISO) and NFW rotation curve fits for NGC 2915. The panels on the left show the fits using the pseudo-isothermal sphere while the panels on the right show the fits using an NFW halo. The top panels show the fits when the stellar rotation curve is scaled by our predetermined value of $\Upsilon_*=0.3$ while the middle and bottom panels shows the fits for the maximum and minimum disk cases respectively. In all panels the black filled circles represent the observed rotation curve while the error bars represent the r.m.s. spread of velocities in a given ring. The black dot-dash curve shows the rotation curve of the gas while the black dotted curve shows the rotation curve of the stellar disk. The solid grey curve shows the resulting rotation curve of the DM halo. Finally, the solid black curve represents the total rotation curve resulting from the best-fit model. In each panel the $\chi^2$ goodness-of-fit statistic is presented. The parameter radial profiles used to constructed the stellar and gas rotation curves were those of tilted ring model SI. The observed rotation curve is also from tilted ring model SI.} \label{rotmas_TR_A_plot} \end{centering} \end{figure*} The fitted parameters for all of the mass models are summarised in Table \ref{mass_modeling_results}. The results for mass model SI are also presented in Fig. \ref{rotmas_TR_A_plot}. This figure shows that NGC 2915 is dark-matter-dominated at nearly all radii (depending on modeling assumptions). We calculate mass ratios as well as the total mass to $B$-band light ratio at the last measured point of each fitted total rotation curve. For mass models SI and CI, DM constitutes, on average, 92.8$\%$ of the dynamical mass. This dark-matter-dominance is confirmed by the total-mass-to-light ratio of NGC 2915 which is 141 $M_{\odot}$/$L_{\odot}$ and 129 $M_{\odot}$/$L_{\odot}$ on average for mass models SI and CI respectively. These estimates are almost double that of the 76~$M_{\odot}$/$L_{\odot}$ upper limit set by \citet{meurer2} thereby making NGC~2915 one of the most dark-matter-dominated late-type dwarf galaxies known. The second main mass-modeling result that Fig. \ref{rotmas_TR_A_plot} demonstrates is that, for either parameterisation of the DM halo, the steeply rising portion of the observed rotation curve is poorly matched by the total rotation curve. It is usually the stellar rotation curve that is used to match the inner observed rotation curve \citep{swaters_thesis}, yet no such scaling is useful in the case of NGC 2915 due to the intrinsically different shapes of the stellar and observed rotation curves. The stellar disk is clearly contained well within the steeply rising portion of the observed rotation curve which continues to rise as $V(R)\propto R$ out to $R\sim 150''$. The best-fitting mass models are those in which the stellar rotation curve contributes zero to the observed total rotation curve. Figure \ref{rotmas_TR_A_plot} shows that when the fitting routine attempts to fit the inner observed rotation curve with the DM rotation curve, the result is an over-estimation of the observed rotation velocities for $R\lesssim 2$ kpc (110$''$) and an under-estimation for radii 2~kpc~$\lesssim R \lesssim$~5.5~kpc. The poor matches between the total and observed inner rotation curves do not allow a particular DM halo parameterisation to be confidently ruled out or confirmed. Furthermore, the observed non-Gaussian line profiles near the centre of NGC 2915 (Sec. \ref{line_profiles}) lead to uncertainties in the shape of the inner portion of the observed rotation curve. These, in turn, lead to associated uncertainties in the shapes of the inner portions of the fitted DM halos. Assuming that the stars are less susceptible than the gas to the effects of non-cricular motions, spectroscopic observations of the stars would have to be carried out to better constrain the inner rotation curve. What is the explanation for the discrepancy between the observed and total rotation curves? Have we failed to include a significant mass component during the decomposition process? A stellar bulge cannot be the answer since our derived 3.6~\micron\ surface brightness profile shows no significant deviation from an exponential light distribution. A significant mass of molecular gas at inner radii is a possibility, yet CO emission in low-mass galaxies is usually weak or not detected \citep{taylor_1998,leroy_2005}. However, the conversion factor used to convert CO luminosity to molecular gas mass is largely uncertain If we consider the mass-modeling results as they appear in Table \ref{mass_modeling_results} then the pseudo-isothermal sphere allows for the best match to the observed rotation curve. The fitted pseudo-isothermal sphere, under the minimum disk assumption, has a core density and a core radius of $\rho_0=0.17\pm~0.03$~$M_{\odot}$~pc$^{-3}$ and $r_c=0.9\pm 0.1$~kpc for model SI. These results are similar to those of \citet{meurer2} who, for their favoured mass model, estimated $\rho_0=0.10\pm~0.02$ ~$M_{\odot}$~pc$^{-3}$ and $r_c=1.23\pm 0.15$~kpc. We also fitted the mass models with $\Upsilon_*$ as a free parameter yet always found the best-fitting value to be negative, consistent with the mass modeling results of \citet{meurer2}. In Fig. \ref{THINGSmm} we compare our derived pseudo-isothermal halo parameters for NGC 2915 to those derived for the THINGS sample from \citet{THINGS_deblok}. The infrared stellar mass-to-light ratios used in their mass models were derived from the 3.6~\micron\ images in combination with stellar population synthesis arguments. Our models for the minimum disk and $\Upsilon_*=0.3$ cases for both models SI and CI are shown in Fig. \ref{THINGSmm}. This comparison suggests that NGC 2915 has a central DM core that is very compact and dense compared to other late-type systems. \begin{figure} \begin{centering} \includegraphics[angle=0,width=1\columnwidth]{THINGS_mm_paperV2.eps} \caption{Mass modeling results for the minimum disk and $\Upsilon_*=0.3$ cases in which the DM halo is modelled as a pseudo-isothermal sphere, compared to the mass modeling results of the THINGS sample of \citet{THINGS_deblok}. The black-filled and open triangles correspond to the minimum disk results for models SI and CI respectively while the black-filled and open squares correspond to the $\Upsilon_*=0.3$ results for models SI and CI respectively. The mass modeling results from THINGS are shown as open circles for dwarf galaxies and filled circles for all other galaxies.} \label{THINGSmm} \end{centering} \end{figure} \section{Summary and conclusions}\label{conclusions} We have obtained new deep, high-resolution H\,{\sc i}\ synthesis observations of NGC~2915 and have used them to carry a detailed study of the H\,{\sc i}\ distribution and dynamics of this galaxy. The galaxy has very different optical and H\,{\sc i}\ properties. Optically, it is similar to other dwarf galaxies, consisting of a small stellar disk. Surrounding the stellar disk, however, is a huge H\,{\sc i}\ disk extending out to galactocentric radii greater than $R= 510''$ with well-defined spiral structure in its outer parts. We estimated the total mass of the H\,{\sc i}\ disk to be $M_{HI}\sim 4.4 \times 10^8$ $M_{\odot}$. Our high resolution, high sensitivity observations clearly resolve the inner H\,{\sc i}\ distribution of NGC~2915 into two H\,{\sc i}\ concentrations separated by $\sim$ 1.1 kpc. This observation discredits the claim by previous investigators that the central gas disk of the system is made up of a massive H\,{\sc i}\ bar. We have shown the shapes of H\,{\sc i}\ line profiles to be significantly non-Gaussian in the vicinity of these central H\,{\sc i}\ over-densities, being split by $\sim$~30~km~s$^{-1}$\ on average over the spatial extent of the lower H\,{\sc i}\ concentration. These profiles are indicative of non-circular velocity components within the gas near the centre of NGC~2915. In our forthcoming paper (Elson et al., 2010, in prep.) we investigate in further detail the central gas dynamics of this system. Linked to these non-circular velocity components, we have provided speculative evidence for the possibility of cold gas accretion from the inter-galactic medium onto the outer disk of NGC~2915 which, if confirmed, would have significant implications for the evolutionary history of this system. These complicated dynamics lead to the new picture that NGC~2915 is not a simple, isolated, low-surface-brightness galaxy but rather an evolving system with a complex interplay between its various mass components. The main focus of this paper has been the regular H\,{\sc i}\ dynamics of NGC~2915. We have fitted a tilted ring model to the H\,{\sc i}\ velocity field. The best-fitting model was one in which the H\,{\sc i}\ disk is severely warped, with the inner and outer disks inclined at $i\sim$ 70 - 75 $^{\circ}$\ and $i \sim$ 55$^{\circ}$\ respectively. This result is, however, made uncertain by the non-Gaussian line profiles near the centre of the galaxy. We therefore also fitted a model with an almost constant inclination of $i\sim$~55$^{\circ}$\ for the entire H\,{\sc i}\ disk. Both models yielded a rotation curve typical of a late-type spiral with $V(R)\propto R$ rotation out to $R\sim 150''$ and constant velocity thereafter. Rotation curves separately derived for the approaching and receding sides of the galaxy differed significantly within their steeply rising portions, providing the clear evidence that NGC~2915 is a kinematically lopsided galaxy. Our results show that these rotation curves cannot be accounted for by the stellar or gas disks of NGC~2915. By using the observed rotation curves as mass model inputs, we found the galaxy to be dark-matter-dominated at nearly all radii with the stellar disk unable to account for the $V(R)\propto R$ portion of the rotation curve. The best-fitting mass model is one in which the stellar disk does not contribute to the observed rotation curve and in which the inferred dark matter halo is parameterised as a pseudo-isothermal sphere. This model has a central core density of $\rho_0= 0.17 \pm 0.03$~$M_{\odot}$~pc$^{-3}$ and a core radius of $R_c = 0.9 \pm 0.1$~kpc. The non-circular gas motions at inner radii lead to uncertainties in the fitted dark matter halo parameters. These, together with the fact that neither DM halo parameterisation allows for an accurate match of the observed rotation curve at inner radii, do not allow a particular DM halo parameterisation to be confidently ruled out or confirmed. Further observations and modeling are required to more accurately determine the inner rotation curve of NGC~2915 and to therefore more confidently discriminate between various parameterisations of the DM halo. It is clear, however, that NGC~2915 has a very dense and compact dark matter core. At the last measured point on the rotation curve, this galaxy has a total-mass-to-light ratio of $M/L\sim$~140~$M_{\odot}$/L$_{\odot}$, almost twice that of the upper limit placed my \citet{meurer2}, thereby making it one of the most dark-matter-dominated galaxies known. \section{Acknowledgments} The work of ECE is based upon research generously supported by the South African SKA project. ECE would like to thank Prof. Renzo Sancisi for his enthusiasm and endless patience in discussing and helping to analyse the H\,{\sc i}\ data. All authors acknowledge funding recieved from the South African National Research Foundation. The work of WJGDB is based upon research supported by the South African Research Chairs Initiative of the Department of Science and Technology and the National Research Foundation. The Australian Telescope Compact Array is part of the Australia Telescope which is funded by the Commonwealth of Australia for operation as a National Facility managed by CSIRO. This work is based [in part] on observations made with the \emph{Spitzer} Space Telescope, which is operated by the Jet Propulsion Laboratory, California Institute of Technology under a contract with NASA. Finally, all authors thank the referee, Claude Carignan, for constructive comments that improved the quality of the paper.
\section{Introduction} Isoscalar giant resonances have been extensively studied since the discovery of the isoscalar giant quadrupole resonance (ISGQR) in the early 1970s \cite{rp1,futo,lebe}. The isoscalar giant monopole resonance (ISGMR) was identified in 1977 \cite{mnh2,dhy2} and was the subject of a number of studies through the 1980s \cite{av1,dhy1,sharma}. The isoscalar giant dipole resonance (ISGDR) was first reported by Morsch {\it {\em et al.}} \cite{hpm1} in $^{208}$Pb but was conclusively identified by Davis {\it {\em et al.}} \cite{bfd1}. Both ISGMR and ISGDR are classified as compression modes and provide information about nuclear incompressibility, $K_{A}$, from which the incompressibility of infinite nuclear matter, K$_\infty$, may be obtained \cite{th-JPB-1}. Most of the earlier investigations of the isoscalar giant resonances used inelastic $\alpha$ scattering at 100--200 MeV and the strength of a particular giant resonance was assumed to be concentrated in a single peak with a Gaussian or Lorentzian shape. The resonance parameters were obtained by multiple-peak fits to the inelastic scattering spectra, after subtraction of a suitable ``background'' ~\cite{speth,mnh1}. In the last decade, the Texas A\&M (TAMU) group has carried out ($\alpha, \alpha'$) studies of many nuclei at a bombarding energy of 240 MeV and extracted the strength distributions of various isoscalar giant resonances in a number of nuclei \cite{exp-TAM-8,exp-TAM-9,exp-TAM-7,exp-TAM-1, john,exp-TAM-3,exp-TAM-2,exp-TAM-6,exp-TAM-5,exp-TAM-11,exp-TAM-10} using a multipole decomposition analysis (MDA)~\cite{exp-bb-1}. Contemporaneously, we have carried out giant resonance measurements using inelastic scattering of 386 MeV $\alpha$ particles at extremely small angles, including 0$^{\circ}$ \cite{ugberk,exp-RCNP-3,exp-RCNP-2,exp-RCNP-5,exp-RCNP-1,exp-ND-3,itoh,exp-ND-2,tao,ug7}. An especially useful feature of our measurements has been the elimination of all instrumental background events from the inelastic scattering spectra which was rendered possible by the optical properties of the Grand Raiden spectrometer~\cite{exp-RCNP-6}. Here, we report on measurements of the isoscalar giant resonances in the even-A Sn isotopes (A=112--124) over the excitation-energy range 8.5--31.5 MeV. Previously, giant resonance measurements on the Sn isotopes have been reported by the TAMU group \cite{dhy1,dhy4} and KVI group \cite{sharma}, using inelastic $\alpha$ scattering at 120--130 MeV and peak-fitting analyses of spectra. More recently, the strength distributions of various isoscalar resonances have been obtained in some Sn isotopes (A=112,116,124) by the TAMU group \cite{exp-TAM-3,exp-TAM-6}. High-quality measurements of the ISGMR over the full range of isotopes provide the opportunity to investigate the asymmetry term, $K_{\tau}$, of the nuclear incompressibility. This term, associated with the neutron excess ($N - Z$), is crucial in obtaining the radii of neutron stars in equation-of-state (EOS) calculations~\cite{exp-JML-1,latti,stein,exp-BAL-2}; the asymmetry ratio, [$(N - Z)/A$], changes by more than 80\% over the range of the investigated Sn isotopes. \section{Experimental Techniques} The experiment was performed at the ring cyclotron facility of the Research Center for Nuclear Physics (RCNP), Osaka University, using inelastic scattering of 386-MeV $\alpha$ particles at extremely forward angles, including 0$^\circ$. A $^4$He$^{++}$ beam was accelerated by the Azimuthally Varying Field (AVF) cyclotron up to 86 MeV, injected into the K = 400 ring cyclotron for acceleration up to 386 MeV, and achromatically transported to the WS experimental hall without any defining slits. To reduce the background at and near 0$^{\circ}$, the beam halo has to be tuned carefully in the experiment. This was accomplished by tuning the beam profile of the injection beam for the ring cyclotron, and typically less than 1 out of 10000 events had contamination from the other bunches. The beam current was 1-20 nA, which was limited by the data acquisition rate and by the maximum available current of the accelerator. The energy resolution obtained was $\sim150$ keV full width at half maximum (FWHM). Self supporting target foils of enriched even-A $^{112-124}$Sn isotopes of thickness 5.0--9.25 mg/cm$^2$ were employed; we used special target frames with a large aperture in order to reduce the background caused by the beam-halo hitting the frames. Data were also taken with a $^{nat}$C target at the actual field settings used in the experiments and the energy calibration was obtained from the peak positions of the 7.652- and 9.641-MeV states in the $^{12}$C$(\alpha,\alpha')$ spectra. Inelastically-scattered $\alpha$ particles were momentum-analyzed with the high-resolution magnetic spectrometer, ``Grand Raiden'' \cite{exp-RCNP-6}, and the vertical and horizontal positions of the $\alpha$ particles were measured with the focal-plane detector system comprised of two position-sensitive multi-wire drift chambers (MWDCs) and two scintillators \cite{exp-RCNP-5}. The MWDCs and scintillators enabled us to make particle identification and to reconstruct the trajectories of the scattered particles. The scattering angle at the target and the momentum of the scattered particles were determined by the ray-tracing method. The vertical-position spectrum obtained in the double-focusing mode of the spectrometer was exploited to eliminate the instrumental background \cite{exp-RCNP-2, exp-RCNP-5}. Giant-resonance data were taken with the spectrometer central angle ($\theta_{spec}$) set at 0$^\circ$, 2.5$^\circ$, 3.5$^\circ$, 5.0$^\circ$, 6.5$^\circ$, and 8.0$^\circ$, covering the angular range from 0$^\circ$ to 9.0$^\circ$ in the laboratory system. The actual angular resolution of the MWDCs, including the nominal broadening of scattering angle due to the emittance of the $^4$He$^{++}$ beam and the multiple Coulomb-scattering effects, was about 0.15$^\circ$ \cite{itohthesis}. The vertical acceptance was limited to $\pm$20 mr by a 2-mm-thick tantalum collimator. The energy bite of the Grand Raiden spectrometer and the special MWDC arrangements for the 0$^{\circ}$ measurements limited the excitation energy range observed to $E_x$=8.5--31.5 MeV. The incident $^4{\rm He}^{++}$ beam was stopped at three independent Faraday Cups (FC) according to the different settings of Grand Raiden \cite{exp-RCNP-5}. In the measurements at large angles ($\theta \geq 6.5^{\circ}$), the beam was stopped on the FC mounted inside the scattering chamber (SC-FC). For measurements at $0^{\circ}$, the FC was located downstream of the MWDCs \cite{exp-RCNP-5}; the incident $^4$He$^{++}$ particles were guided to a vacuum pipe situated at the high-momentum side of the MWDCs and finally stopped at the $0^{\circ}$-FC. A third FC was used for measurements in the scattering-angle region $2^{\circ} \leq \theta \leq 5^{\circ}$. This FC was installed behind the Q1-magnet of the Grand Raiden (Q1-FC)~\cite{itoh-rep}. The use of these three Faraday-cups allowed us to obtain reliable values of accumulated charges for the incident $^4{\rm He}^{++}$ beam at different scattering angles. Normalization of the FCs was obtained with a beamline polarimeter located upstream from the target. The polarimeter target was inserted in the beam, and the scattered $^{4}$He$^{++}$ particles counted, before and after each change of the FC. The overall accuracy of this normalization is estimated to be $\sim$2\%, including systematic errors from slight changes in the direction of the beam during the measurement \cite{itohthesis}. The ion-optics of the Grand Raiden spectrometer is such that the particles scattered from the target position are focused vertically and horizontally at the focal plane. Using this property, the instrumental background was completely eliminated. While inelastically scattered $\alpha$ particles are focused at the focal plane, background events due to the rescattering of $\alpha$ particles from the wall and pole surfaces of the spectrometer show a flat distribution in the vertical-position spectra at the focal plane, as shown in Fig.~\ref{fig:ypos}. The peak region in the vertical position spectrum was treated as true+background events. The off-center regions were treated as background only. Figure~\ref{fig:xpos}(a) shows the horizontal position spectrum for the $^{112}$Sn($\alpha,\alpha'$) reaction at 0$^{\circ}$. The background spectrum has no distinct structure in the giant resonance region. Finally, we have obtained clean spectra by subtracting the background spectrum from the true+background spectrum, as shown in Fig.~\ref{fig:xpos}(b). \begin{figure}[htpb] \begin{center} \includegraphics[scale=0.8]{fig1} \caption{\label{fig:ypos} Vertical-position spectrum at the focal plane of the Grand Raiden spectrometer, taken at 2.5$^\circ$. The central densely-hatched region represents true+background events. The off-center sparsely-hatched regions represent only background events. The real events were obtained by subtracting background events from the true+background events.} \end{center} \end{figure} \begin{figure}[htpb] \begin{center} \includegraphics[scale=0.8]{fig2} \caption{\label{fig:xpos} (a) Horizontal-position spectrum of the $^{112}$Sn($\alpha$,$\alpha'$) reaction at 0$^\circ$. The hatched region is background events. (b) Background-free spectrum } \end{center} \end{figure} The background-free ``$0^\circ$'' inelastic spectra for the Sn isotopes are presented in Fig.~\ref{fig:sn0deg}. In all cases, the spectrum is dominated by the ISGMR+ISGQR peak near $E_x\sim15$ MeV. There is an underlying continuum in the high excitation-energy region in the spectrum; it is reasonable to assume that this continuum, remaining after elimination of the instrumental background, is primarily due to contributions from excitation of the higher multipoles and quasifree knockout processes \cite{brand}. \begin{figure}[htpb] \begin{center} \includegraphics[scale=1.1]{fig3} \caption{\label{fig:sn0deg} Excitation-energy spectra obtained from inelastic $\alpha$ scattering at $\theta_{lab}$ = 0.69$^\circ$ for all even-A Sn isotopes.} \end{center} \end{figure} \section{Data Analysis} We have employed the MDA procedure~\cite{exp-bb-1} to extract the strengths of the isoscalar giant monopole resonance (ISGMR), the isoscalar giant dipole resonance (ISGDR), the isoscalar giant quadrupole resonance (ISGQR), and the high-energy octupole resonance (HEOR) in the Sn isotopes. The cross-section data were binned into 1-MeV energy intervals to reduce the statistical fluctuations. For each excitation energy bin from 8.5 MeV to 31.5 MeV, the experimental 17-point angular distribution $\frac{d\sigma^{exp}}{d\Omega}(\theta_{cm},E_x)$ has been fitted by means of the least-square method with the linear combination of calculated distributions $\frac{d\sigma^{cal}_{L}}{d\Omega}(\theta_{cm},E_x)$ defined by: \begin{equation} \frac{d\sigma^{exp}}{d\Omega}(\theta_{cm},E_x)=\sum_{L=0}^{7}a_L(E_x) \times \frac{d\sigma^{cal}_{L}}{d\Omega}(\theta_{cm},E_x) \end{equation} \noindent where $\frac{d\sigma^{cal}_{L}}{d\Omega}(\theta_{cm},E_x)$ is the calculated distorted-wave Born approximation (DWBA) cross section corresponding to 100\% energy-weighted sum rule (EWSR) for the L-th multipole. The fractions of the EWSR, $a_L(E_x)$, for various multipole components were determined by minimizing $\chi^2$. This procedure is justified since the angular distributions are well characterized by the transferred angular momentum $L$, according to the DWBA calculations for $\alpha$ scattering. It was confirmed that the MDA fits were not affected by including $L>$7. The DWBA calculations were performed following the method of Satchler and Khoa \cite{th-GRS-1}, using the density-dependent single-folding model for the real part, obtained with a Gaussian $\alpha$-nucleon potential, and a phenomenological Woods-Saxon potential for the imaginary term. Therefore, the $\alpha$-nucleus interaction is given by: \begin{equation} \label{potential} U(r)=V_F(r)+iW/(1+exp((r-R_I)/a_I)) \end{equation} \noindent where $V_F(R)$ is the real single-folding potential obtained by folding the ground-state density with the density-dependent $\alpha$-nucleon interaction: \begin{equation} v_{DDG}(r,r',\rho)=-v(1-\beta\rho(r')^{2/3})exp(-|r-r'|^2/t^2)) \nonumber \end{equation} \noindent where $v_{DDG}(r,r',\rho)$ is the density-dependent $\alpha$-nucleon interaction, $|r-r'|$ is the distance between the center of mass of the $\alpha$ particle and a target nucleon, $\rho(r')$ is the ground-state density of the target nucleus at the position $r'$ of the target nucleon, $\beta$=1.9 fm$^2$, and $t$=1.88 fm. W is the depth of the Woods-Saxon type imaginary part of the potential, with the reduced radius $R_I$ and diffuseness $a_I$. These calculations were performed with the computer code PTOLEMY~\cite{ptolemy,ptolemy2}, with the input values modified~\cite{satchler} to take into account the correct relativistic kinematics. The shape of the real part of the potential and the form factor for PTOLEMY were obtained using the codes SDOLFIN and DOLFIN \cite{dolfin}. We use the transition densities and sum rules for various multipolarities described in Refs.~\cite{mnh1,satchler2,mnh3}. The radial moments were obtained by numerical integration of the Fermi mass distribution with the parameter values from Ref.~\cite{data-1} (listed in Table~\ref{tab:fermi}). \begin{table} \caption{\label{tab:fermi}Fermi-distribution parameters from Ref.~\cite{data-1}. ``c'' is the adjusted half-density radius for the charge distribution and ``a'' is the diffuseness parameter.} \begin{ruledtabular} \begin{tabular}{cccccccc} Target&$^{112}$Sn&$^{114}$Sn&$^{116}$Sn&$^{118}$Sn&$^{120}$Sn&$^{122}$Sn&$^{124}$Sn \\\hline c$ (fm)$&5.3714&5.3943&5.4173&5.4391&5.4588&5.4761&5.4907 \\ a$ (fm)$&0.523&0.523&0.523&0.523&0.523&0.523&0.523 \\ \end{tabular} \end{ruledtabular} \end{table} \begin{figure}[htpb] \includegraphics[scale=0.8]{fig4 \caption{\label{fig:112sn-elas} (a) Ratio of the elastic $\alpha$-scattering cross sections to the Rutherford cross sections for $^{112}$Sn at 386 MeV. (b) Differential cross sections for excitation of the 2$^+_1$ state in $^{112}$Sn. The solid lines are the results of the folding-model calculations.} \end{figure} \begin{figure}[htpb] \includegraphics[scale=0.8]{fig5 \caption{\label{fig:120sn-elas} Same as Fig.~\ref{fig:112sn-elas}, except for $^{120}$Sn.} \end{figure} \begin{figure}[htpb] \includegraphics[scale=0.8]{fig6 \caption{\label{fig:124sn-elas} Same as Fig.~\ref{fig:112sn-elas}, except for $^{124}$Sn.} \end{figure} The optical-model (OM) parameters {\em viz.} the real part of the potential ($V_F(r)$), the Woods-Saxon type imaginary part of potential ($W$), the reduced radius ($R_I$), and the diffuseness ($a_I$) in Eq.~\ref{potential} were determined by fitting the differential cross sections of elastic $\alpha$ scattering measured for $^{112}$Sn, $^{120}$Sn, and $^{124}$Sn in a companion experiment; the results are listed in Table~\ref{tab:potential}. The OM fits to the elastic scattering data for $^{112}$Sn, $^{120}$Sn, and $^{124}$Sn, are shown in Figs.~\ref{fig:112sn-elas}(a),~\ref{fig:120sn-elas}(a), and \ref{fig:124sn-elas}(a), respectively. To test the efficacy of the OM parameters, DWBA calculations were carried out for the first 2$^+$ states in these nuclei using a collective form factor and previously-established B(E2) values obtained from Refs.~\cite{bel1,bel2} (also listed in Table~\ref{tab:potential}). Figures.~\ref{fig:112sn-elas}(b), ~\ref{fig:120sn-elas}(b), and~\ref{fig:124sn-elas}(b) compare the results of these calculations with the experimental data; indeed, the DWBA calculations reproduce the experimental differential cross sections for the 2$^+_1$ states well without any normalization. \begin{table} \caption{\label{tab:potential}OM parameters obtained by fitting elastic scattering data. Also listed are the B(E2) values for the corresponding 2$^+_1$ states from Refs.~\cite{bel1,bel2}.} \begin{ruledtabular} \begin{tabular}{ccccccc} Target&V (MeV)&W (MeV)&$a_I$ (fm)&$R_I$ (fm)& B(E2) ($e^2b^2$) \\\hline $^{112}$Sn&33.9 & 31.7 & 0.60 & 1.02 &0.24\\ $^{120}$Sn&33.4 & 33.0 & 0.63 & 1.01 & 0.20\\ $^{124}$Sn&34.0 & 33.5 & 0.61 & 1.02 & 0.17\\ \end{tabular} \end{ruledtabular} \end{table} The contribution of the IVGDR excitation to the measured cross sections was subtracted prior to multipole decomposition. Cross sections for exciting the IVGDR were obtained with DWBA calculations on the basis of the Goldhaber-Teller model and using the strength distribution obtained from photonuclear work \cite{Dietrich}. Figs.~\ref{fig:ang} and~\ref{fig:ang2} show the MDA fits to the experimental angular distributions of the differential cross sections for the 16.5-MeV and 25.5-MeV energy bins in the inelastic-scattering spectra of $^{112}$Sn and $^{124}$Sn, respectively, along with the contributions from the $L$=0, 1 and 2 multipoles. The ISGMR contribution is dominant in comparison to the other multipoles at E$_x$=16.5 MeV. On the other hand, the ISGDR is the dominant contributor at E$_x$=25.5 MeV. \begin{figure}[htpb] \includegraphics[scale=1]{fig7} \caption{\label{fig:ang} Angular distribution of 1-MeV bins centered at E$_x$=16.5 MeV for $^{112}$Sn($\alpha$,$\alpha'$) and $^{124}$Sn($\alpha$,$\alpha'$). The solid squares are the experimental data and the solid lines are the MDA fits to the data. Also shown are the contributions to the fits from $L$=0 (dashed line), $L$=1 (dotted line), $L$=2 (dash-dotted line) and $L$=3 (small-dashed line) multipoles, as well as from the IVGDR (dash-dot-dotted line).} \end{figure} \begin{figure}[htpb] \includegraphics[scale=1]{fig8} \caption{\label{fig:ang2} Same as Fig.~\ref{fig:ang} except for the 25.5-MeV energy bin (see text).} \end{figure} \section{Results and Discussion} We have extracted strength distributions for $L$=0, 1, 2, and 3 multipoles over the energy range 8.5 MeV--31.5 MeV in all the Sn isotopes investigated in this work. These are displayed in Figs.~\ref{fig:ISGMR}, \ref{fig:gdr},~\ref{fig:ISGQR}, and~\ref{fig:heor}, respectively. The strengths are related to the coefficients $a_L$ in Eq.~1 as (see, Refs. ~\cite{mnh1,mnh3}): \begin{equation} S_0(E_x)=\frac{2\hbar^2A<r^2>}{mE_x}a_0(E_x) \end{equation} \begin{eqnarray} S_1(E_x)=\frac{3\hbar^2A}{32\pi mE_x}(11<r^4>-\frac{25}{3}<r^2>^2 -10\epsilon<r^2>)a_1(E_x) \end{eqnarray} \begin{equation} S_{L\geq2}(E_x)=\frac{\hbar^2A}{8\pi mE_x}L(2L+1)^2<r^{2L-2}>a_2(E_x) \end{equation} \noindent where $m$, $A$ and $<r^N>$ are the nucleon mass, the mass number, and the $N$th moment of the ground-state density, and $\epsilon$=(4/$E_2$+5/$E_0$)$\hbar^2$/3$mA$; $E_0$ and $E_2$ are the centroid energies of the ISGMR and the ISGQR and have been taken as 80~A$^{-1/3}$ MeV and 64~A$^{-1/3}$ MeV, respectively. It should be noted that although we employed calculated DWBA cross sections with up to $L$=7 in the MDA fitting procedure, it was not possible to reliably extract the strength distributions for $L\geq$4 because of the limited angular range (0$^{\circ}$--9$^{\circ}$). Further, there is a small, near-constant ISGMR and ISGQR strength up to the highest excitation energies measured in this experiment. The {\em raison d`\^{e}tre} of this extra strength is not quite well understood. However, similarly enhanced E1 strengths at high excitation energies were noted previously \cite{exp-RCNP-5,exp-RCNP-1} and have been attributed to contributions to the continuum from three-body channels, such as knockout reactions \cite{brand}. These processes are implicitly included in the MDA as background and may lead to spurious contributions to the extracted multipole strengths at higher energies where the associated cross sections are very small. This conjecture is supported by measurements of proton decay from the ISGDR at backward angles wherein no such spurious strength is observed in spectra in coincidence with the decay protons~\cite{exp-ND-3,hun1,nayak2,hun3}; quasifree knockout results in protons that are forward peaked. A similar increase in the ISGMR strength at high excitation energies was reported as well by the TAMU group in $^{12}$C when they carried out MDA without subtracting the continuum from the excitation-energy spectra~\cite{john}. \begin{figure} \includegraphics[scale=1.4]{fig9} \caption{\label{fig:ISGMR} (Color online) ISGMR strength distributions obtained for the Sn isotopes in the present experiment. Error bars represent the uncertainties from fitting the angular distributions in the MDA procedure. The solid lines show Lorentzian fits to the data.} \end{figure} \begin{figure} \includegraphics[scale=1.4]{fig10} \caption{\label{fig:gdr} (Color online) ISGDR strength distributions obtained for the Sn isotopes in the present experiment. Error bars represent the uncertainties from fitting the angular distributions in the MDA procedure. The solid lines show Lorentzian fits to the data.} \end{figure} \begin{figure} \includegraphics[scale=1.4]{fig11} \caption{\label{fig:ISGQR} (Color online) ISGQR strength distributions obtained for the Sn isotopes in the present experiment. Error bars represent the uncertainties from fitting the angular distributions in the MDA procedure. The solid lines show Lorentzian fits to the data.} \end{figure} \begin{figure}[htbp] \center \includegraphics[scale=1.2]{fig12} \caption{\label{fig:heor} (Color online) HEOR strength distributions obtained for the Sn isotopes in the present experiment. Error bars represent the uncertainties from fitting the angular distributions in the MDA procedure. The solid lines show Lorentzian fits to the data.} \end{figure} The $L=0$ strength distributions were fitted with a Lorentzian function to determine the centroid energies and widths of the ISGMR. These fits are shown superimposed in Fig.~\ref{fig:ISGMR}; the corresponding fitting parameters are presented in Table~\ref{tbl:ISGMRfit} and compared with results from TAMU~\cite{exp-TAM-6, exp-TAM-3}. In this and subsequent comparisons and discussion, we refer only to the recent TAMU results because those are from comparable data and analysis---all other previous results on the Sn isotopes were from peak-fitting analyses of data taken at significantly lower energies. In order to compare with the available theoretical results, various moment ratios for the experimental ISGMR strength distributions have been calculated over the excitation-energy range, $E_x$ = 10.5--20.5 MeV, encompassing the ISGMR peak. The results are listed in Table~\ref{tbl:gmrmom}. The reasons for the difference between the present results and those from TAMU for $^{112}$Sn and $^{124}$Sn are not readily apparent but might be attributable to the fact that in their analysis the multipole decomposition is carried out after subtracting a ``background'' from the excitation-energy spectrum, whereas, as pointed out earlier, no such subtraction is required in the present analysis since the Sn($\alpha,\alpha'$) spectra obtained in our work have been rendered free of all instrumental background events. Figure~\ref{fig:gdr} shows the strength distributions of ISGDR. We observe a ``bi-modal'' distribution between $E_x$=8.5 MeV and $E_x$=31.5 MeV. This bi-modal pattern for the ISGDR has been observed in all nuclei investigated so far, both in the RCNP and TAMU measurements. This ``low-energy'' isoscalar $L$=1 strength (LE) has engendered considerable interest and argument over the past few years. It is present in nearly all of the recent theoretical calculations in some form or the other, and at similar energies, although with varying strength. It has been shown \cite{co00,dv00} that the centroid of this component of the $L$=1 strength is independent of the nuclear incompressibility and while the exact nature of this component is not fully understood yet, suggestions have been extended to the effect that this component might represent the ``toroidal''~\cite{balb,dv00} or the ``vortex'' modes~\cite{dubna}. It is impossible to distinguish between the competing possibilities based on currently-available data~\cite{exp-ND-3}. There is general agreement, however, that only the high-energy (HE) component of this bi-modal distribution needs to be considered in obtaining a value of $K_\infty$ from the energy of the ISGDR. The strength distributions of the ISGDR, therefore, have been fitted with a two-Lorentzian function and the fitting parameters for the LE- and HE-components are presented in Tables~\ref{tbl:gdrfitLE} and~\ref{tbl:gdrfitHE}, respectively. It may be noted that because of the ``spurious'' strength at the higher excitation energies mentioned previously, the numbers for the extracted EWSR are significantly larger than 100\% in some cases. The strength distributions of the ISGQR are shown in Fig.~\ref{fig:ISGQR}. These too were fitted with a Lorentzian function to determine the centroid energies and the widths. The fit parameters are presented in Table~\ref{tbl:ISGQRfit}. The $L$=3 strength distributions (Fig.~\ref{fig:heor}) show an enhanced strength at $E_{x} <$ 10 MeV. This part is, most likely, from the low-energy octupole resonance (LEOR). The LEOR represents the 1$\hbar\omega$ component of the $L$=3 strength and has been reported in a number of nuclei previously \cite{moss1,moss2}. The strength distributions were, therefore, fitted with a two-Lorentzian function to determine the centroid energy of HEOR (the high-excitation-energy component). The extracted HEOR peak-energies are presented in Table~\ref{tbl:heorfit}. \begin{table}[tpb] \caption{\label{tbl:ISGMRfit}Lorentzian-fit parameters for the ISGMR strength distributions in the Sn isotopes, as extracted from MDA. The quoted EWSR values are from the fitted Lorentzians. The results from TAMU work (from Gaussian fits), where available, are provided for comparison~\cite{exp-TAM-6,exp-TAM-3}.} \begin{ruledtabular} \begin{tabular}{lrrrr} \multicolumn{1}{c}{Target} & $E_{ISGMR}$ (MeV)&$\Gamma$ (MeV)& EWSR\footnotemark[1]& Reference\\ \hline $^{112}$Sn & $16.1\pm0.1$ & $4.0\pm0.4$ & $0.92\pm0.04$ & This work\\ & $15.67^{+0.11}_{-0.11}$& $5.18^{+0.40}_{-0.04}$ & $1.10^{+0.15}_{-0.12}$ & TAMU\\ $^{114}$Sn & $15.9\pm0.1$&$4.1\pm0.4$&$1.04\pm0.06$& This work\\ $^{116}$Sn & $15.8\pm0.1$&$4.1\pm0.3$&$0.99\pm0.05$& This work\\ $^{118}$Sn & $15.6\pm0.1$&$4.3\pm0.4$&$0.95\pm0.05$& This work\\ $^{120}$Sn & $15.4\pm0.2$&$4.9\pm0.5$&$1.08\pm0.07$& This work\\ $^{122}$Sn & $15.0\pm0.2$&$4.4\pm0.4$&$1.06\pm0.05$ & This work\\ $^{124}$Sn & $14.8\pm0.2$&$4.5\pm0.5$&$1.05\pm0.06$ & This work\\ & $15.34^{+0.13}_{-0.13}$ & $5.00^{+0.13}_{-0.53}$ & $1.06^{+0.10}_{-0.20}$ & TAMU\\ \end{tabular} \end{ruledtabular} \footnotetext[1] { Only statistical uncertainties are included; systematic errors, mostly from DWBA calculations, are $\sim$15\%.} \end{table} \begin{table}[tpb] \caption{ \label{tbl:gmrmom}Various moment ratios for the ISGMR strength distributions in the Sn isotopes. All moments have been calculated over $E_x$ = 10.5--20.5 MeV. The quoted EWSR values are from the strength observed within this energy range. The results from TAMU work, where available, are provided for comparison~\cite{exp-TAM-6,exp-TAM-3}.} \begin{ruledtabular} \begin{tabular}{lrrrrr} \multicolumn{1}{c}{Target} & $\frac{m_1}{m_0}$ (MeV)&$\sqrt{\frac{m_3}{m_1}}$ (MeV)&$\sqrt{\frac{m_1}{m_{-1}}}$ (MeV)& EWSR\footnotemark[1] & Reference\\ \hline $^{112}$Sn & $16.2\pm0.1$&$16.7\pm0.2$&$16.1\pm0.1$ & $0.73\pm0.04$ & This work\\ & $15.43^{+0.11}_{-0.10}$& $16.05^{+0.26}_{-0.14}$ & $15.23^{+0.10}_{-0.10}$ & $1.16^{+0.13}_{-0.18}$ & TAMU\\ $^{114}$Sn & $16.1\pm0.1$ & $16.5\pm0.2$ & $15.9\pm0.1$ & $0.86\pm0.05$ & This work\\ $^{116}$Sn & $15.8\pm0.1$&$16.3\pm0.2$&$15.7\pm0.1$ & $0.86\pm0.05$& This work\\ & $15.85\pm0.20$& & & $1.12\pm0.15$& TAMU\\ $^{118}$Sn & $15.8\pm0.1$&$16.3\pm0.1$&$15.6\pm0.1$ & $0.73\pm0.04$& This work\\ $^{120}$Sn & $15.7\pm0.1$&$16.2\pm0.2$&$15.5\pm0.1$ & $0.78\pm0.05$& This work\\ $^{122}$Sn & $15.4\pm0.1$&$15.9\pm0.2$&$15.2\pm0.1$ & $0.85\pm0.05$ & This work\\ $^{124}$Sn & $15.3\pm0.1$&$15.8\pm0.1$&$15.1\pm0.1$ & $0.77\pm0.05$ & This work\\ & $14.50^{+0.14}_{-0.14}$ & $14.96^{+0.10}_{-0.11}$ &$14.33^{+0.17}_{-0.14}$ & $1.04^{+0.11}_{-0.11}$ & TAMU\\ \end{tabular} \end{ruledtabular} \footnotetext[1] { Only statistical uncertainties are included; systematic errors, mostly from DWBA calculations, are $\sim$15\%.} \end{table} \begin{table}[tpb] \caption{ \label{tbl:gdrfitLE}Lorentzian-fit parameters for the low-energy component of ISGDR strength distributions in the Sn isotopes, as extracted from MDA. The results from TAMU work, where available, are provided for comparison~\cite{exp-TAM-6,exp-TAM-3}.} \begin{ruledtabular} \begin{tabular}{lrrr} \multicolumn{1}{c}{Target} & $E_{LE-ISGDR}$ (MeV)&$\Gamma$ (MeV)& Reference\\ \hline $^{112}$Sn & $15.4\pm0.1$ & $4.9\pm0.5$ & This work\\ & $14.92^{+0.15}_{-0.14}$& $8.82^{+0.26}_{-0.29}$ & TAMU\\ $^{114}$Sn & $15.0\pm0.1$&$5.6\pm0.5$& This work\\ $^{116}$Sn & $14.9\pm0.1$&$5.9\pm0.5$& This work\\ & $14.38\pm0.25$& $5.84\pm0.30$ & TAMU\\ $^{118}$Sn & $14.8\pm0.1$&$6.1\pm0.3$& This work\\ $^{120}$Sn & $14.7\pm0.1$&$5.9\pm0.3$& This work\\ $^{122}$Sn & $14.4\pm0.1$&$6.7\pm0.3$& This work\\ $^{124}$Sn & $14.3\pm0.1$&$6.6\pm0.3$ & This work\\ & $13.31^{+0.15}_{-0.15}$ & $6.60^{+0.15}_{-0.13}$ & TAMU\\ \end{tabular} \end{ruledtabular} \end{table} \begin{table}[tpb] \caption{ \label{tbl:gdrfitHE}Lorentzian-fit parameters for the high-energy component of ISGDR strength distributions in the Sn isotopes, as extracted from MDA. The results from TAMU work, where available, are provided for comparison~\cite{exp-TAM-6,exp-TAM-3}.} \begin{ruledtabular} \begin{tabular}{lrrrr} \multicolumn{1}{c}{Target} & $E_{HE-ISGDR}$ (MeV)&$\Gamma$ (MeV) & EWSR\footnotemark[1] &Reference\\ \hline $^{112}$Sn & $26.2\pm0.8$ & $16.3\pm4.0$ &$1.02\pm0.03$ & This work\\ & $26.28^{+0.32}_{-0.23}$& $10.82^{+0.39}_{-0.36}$ & $0.70^{+0.10}_{-0.10}$& TAMU\\ $^{114}$Sn & $26.1\pm0.8$&$13.9\pm3.4$& $1.23\pm0.03$& This work\\ $^{116}$Sn & $25.9\pm0.6$&$13.1\pm4.2$& $1.02\pm0.03$&This work\\ & $25.50\pm0.60$& $12.00\pm0.60$ & $0.88\pm0.20$& TAMU\\ $^{118}$Sn & $26.0\pm0.3$&$13.1\pm2.0$& $1.20\pm0.03$& This work\\ $^{120}$Sn & $26.0\pm0.4$&$13.1\pm1.9$&$1.50\pm0.03$& This work\\ $^{122}$Sn & $26.3\pm0.2$&$12.4\pm1.1$ & $1.47\pm0.03$&This work\\ $^{124}$Sn & $25.7\pm0.5$&$10.2\pm1.6$ & $1.29\pm0.06$&This work\\ & $25.06^{+0.22}_{-0.21}$ & $13.87^{+0.28}_{-0.22}$ & $0.93^{+0.12}_{-0.13}$& TAMU\\ \end{tabular} \end{ruledtabular} \footnotetext[1] { Only statistical uncertainties are included; systematic errors, mostly from DWBA calculations and the contributions at the highest energies (see text), are $\sim$30\%.} \end{table} \begin{table}[tpb] \caption{ \label{tbl:ISGQRfit}Lorentzian-fit parameters of ISGQR strength distributions in the Sn isotopes, as extracted from MDA. The results from TAMU work, where available, are provided for comparison~\cite{exp-TAM-6,exp-TAM-3}.} \begin{ruledtabular} \begin{tabular}{lrrrr} \multicolumn{1}{c}{Target} & $E_{ISGQR}$ (MeV)&$\Gamma$ (MeV)& EWSR\footnotemark[1]&Reference\\ \hline $^{112}$Sn & $13.4\pm0.1$ & $7.0\pm0.5$ &$1.08\pm0.04$ & This work\\ & $13.48^{+0.15}_{-0.14}$& $4.90^{+0.22}_{-0.27}$ &$0.88^{+0.14}_{-0.13}$& TAMU\\ $^{114}$Sn & $13.2\pm0.1$&$6.8\pm0.4$& $1.25\pm0.05$ & This work\\ $^{116}$Sn & $13.1\pm0.1$&$6.4\pm0.4$& $1.12\pm0.04$ &This work\\ & $13.50\pm0.35$& $5.00\pm0.30$ & $1.08\pm0.12$& TAMU\\ $^{118}$Sn & $13.1\pm0.1$&$6.6\pm0.3$& $1.08\pm0.03$& This work\\ $^{120}$Sn & $12.9\pm0.1$&$7.0\pm0.7$& $1.04\pm0.04$& This work\\ $^{122}$Sn & $12.8\pm0.1$&$7.8\pm0.6$& $1.25\pm0.04$& This work\\ $^{124}$Sn & $12.6\pm0.1$&$7.7\pm0.9$& $1.13\pm0.04$& This work\\ & $12.72^{+0.11}_{-0.11}$ & $4.20^{+0.32}_{-0.03}$ & $0.89^{+0.15}_{-0.10}$& TAMU\\ \end{tabular} \end{ruledtabular} \footnotetext[1] { Only statistical uncertainties are included; systematic errors, mostly from DWBA calculations, are $\sim$20\%.} \end{table} \begin{table}[tpb] \caption{ \label{tbl:heorfit}Lorentzian-fit parameters of HEOR strength distributions in the Sn isotopes, as extracted from MDA. The results from TAMU work, where available, are provided for comparison~\cite{exp-TAM-6,exp-TAM-3}.} \begin{ruledtabular} \begin{tabular}{lrrr} \multicolumn{1}{c}{Target} & $E_{HEOR}$ (MeV)&$\Gamma$ (MeV)& Reference\\ \hline $^{112}$Sn & $22.7\pm0.7$ & $7.2\pm1.9$ & This work\\ & $20.63^{+0.30}_{-0.28}$& $3.21^{+0.30}_{-0.28}$& TAMU\footnotemark[1]\\ $^{114}$Sn & $22.7\pm0.7$&$7.2\pm2.1$& This work\\ $^{116}$Sn & $22.3\pm0.6$&$7.6\pm1.7$& This work\\ & $23.3\pm0.8$& $10.9\pm0.6$ & TAMU\\ $^{118}$Sn & $22.1\pm0.6$&$5.9\pm1.5$& This work\\ $^{120}$Sn & $22.3\pm0.6$&$7.5\pm1.8$& This work\\ $^{122}$Sn & $22.1\pm0.6$&$5.6\pm1.5$& This work\\ $^{124}$Sn & $22.1\pm0.5$&$8.1\pm1.5$& This work\\ & $19.12^{+0.26}_{-0.26}$& $3.30^{+0.17}_{-0.05}$& TAMU\footnotemark[1]\\ \end{tabular} \end{ruledtabular} \footnotetext[1] {($m_1/m_0$) ratios.} \end{table} \begin{figure}[htbp] \includegraphics[scale=0.4]{fig13} \caption{\label{fig:snall} (Color online) Systematics of the moment ratios $m_1/m_0$ for the ISGMR strength distributions in the Sn isotopes. The experimental results (filled squares) are compared with results from non-relativistic RPA calculations (without pairing) by Col\`o {\em et al.}~\cite{th-GC-3,th-GC-4} (filled circles); relativistic calculations of Piekarewicz~\cite{th-JP-4} (triangles); RMF calculations from Vretenar {\em et al.}~\cite{dario2} (diamonds); and, QTBA calculations from the J\"{u}lich group~\cite{julich} (sideways triangles). Results for $^{112}$Sn, $^{116}$Sn and $^{124}$Sn reported by the TAMU group~\cite{exp-TAM-6,exp-TAM-3} are also shown (inverted triangles). } \end{figure} The primary focus of this work has been on the ISGMR because of its direct connection with the nuclear incompressibility. The excitation energy of the ISGMR is expressed in the scaling model \cite{th-SS-4} as: \begin{equation} \label{eqn:gmr} E_{ISGMR}=\hbar\sqrt{\frac{K_A}{m<r^2>}} \end{equation} \noindent where $m$ is the nucleon mass, $<r^2>$ the ground-state mean-square radius, and $K_{A}$, the incompressibility of the nucleus. The moment ratios, $m_1/m_0$, for the ISGMR strengths in the Sn isotopes are shown in Fig.~\ref{fig:snall} and compared with recent theoretical results from Col\`o {\em et al.} (non-relativistic) \cite{th-GC-3,th-GC-4} and Piekarewicz (relativistic)~\cite{th-JP-4}. The calculations overestimate the experimental ISGMR energies significantly (by almost 1 MeV in case of the higher-A isotopes!). This difference is very surprising since the interactions used in these calculations are those that very closely reproduce the ISGMR energies in $^{208}$Pb and $^{90}$Zr. Indeed, this disagreement leaves open a puzzling question: Why are the tin isotopes so soft \cite{th-JP-4}? Are there any nuclear structure effects that need to be taken into account to describe the ISGMR energies in the Sn isotopes? Or, more provocatively, do the ISGMR energies depend on something more than the nuclear incompressibility, requiring a modification of the scaling relationship given in Eq.~\ref{eqn:gmr}? There have been several attempts to explain this anomaly. One of the earliest was by Civitarese {\em et al.}~\cite{civi} to estimate the effect of pairing correlations on the energy of the ISGMR. The shifts obtained for the ISGMR energies of 100--150 keV across the Sn isotopic chain were insufficient to explain the experimental data. Piekarewicz and Centelles \cite{jorge4} have constructed a hybrid model having the same incompressibility coefficient ($K_{\infty}$=230 MeV) as the FSUGold \cite{th-JP-3} while preserving the stiff symmetry energy of NL3 \cite{lala}. This results in a considerably softer incompressibility coefficient for neutron-rich matter and produces a significant improvement in agreement with the experimental data on the ISGMR's in the Sn isotopes. However, as the authors point out, while the improvement in case of the Sn isotopes is unquestionable, an important problem remains: the hybrid model underestimates the ISGMR centroid energy in $^{208}$Pb by almost 1 MeV, suggesting that the rapid softening with neutron excess predicted by this hybrid model might be unrealistic. They also suggest that the failure of the FSUGold to reproduce the ISGMR energies might be due to missing physics unrelated to the incompressibility of neutron-rich nuclear matter; as an example of such missing physics, they mention the superfluid character of the Sn isotopes resulting from their open-shell structure. Calculations have also become available recently from the RMF approach with the DD-ME2 interaction \cite{dario1}, and these reproduce the centroids of the ISGMR in the Sn isotopes rather well \cite{dario2}. It is also seen that the DD-ME2 interaction falls within the constraints imposed by the experimental $K_{\infty}$ and $K_{\tau}$ values (see discussion below). Some concern has been expressed, however, that this agreement of the centroid energies might be just a coincidence since the ISGMR strength distributions for the Sn isotopes from this work appear to be not significantly different from those obtained from, for example, the FSUGold~\cite{jorgen}. In calculations using the T5 Skyrme interaction within the quasiparticle time blocking approximation (QTBA) approach, Tselyaev {\em et al.} \cite{julich} have obtained the ISGMR strength distributions in all the Sn isotopes in good agreement with the experimental data, including the resonance widths. However, T5 has the associated $K_{\infty}$ value of only 202 MeV, which is significantly lower than that extracted earlier from the ISGMR's in $^{208}$Pb and $^{90}$Zr. While the agreement with the experimental data is impressive (and, indeed, reproduces the A-dependence rather well), it does leave the question of ``softness'' of the Sn nuclei unanswered. As the authors themselves state, the goal of their work has not been to solve the problem of the nuclear-matter incompressibility but to find under which conditions one can obtain reasonable description of the experimental data for the considered tin isotopes. The ``superfluid'' character of the Sn isotopes, resulting from pairing correlations in open-shell nuclei, has been investigated by Li {\em et al.} \cite{colo3}. In a self-consistent QRPA model that employs the canonical HFB basis and an energy-density functional with a Skyrme mean-field part and density-dependent pairing, they calculated the energy of the ISGMR for the Sn isotopes and looked at the effects of different kinds of pairing forces (volume, surface, and mixed). They find that, compared with the HF+RPA and HF-BCS-QRPA formalisms, the HFB+QRPA calculations lead to energies for the ISGMR in Sn isotopes that are significantly closer to the experimental values, in particular with the surface pairing forces and the SKM* interaction ($K_{\infty}\sim$ 215 MeV) \cite{skm}. Thus, while pairing effects lower the ISGMR excitation energies, one still needs to reduce the $K_{\infty}$ value by $\sim$10\% for achieving a reasonable agreement with the experimental data. A very intriguing possibility in explaining the ``softness'' of the Sn isotopes has been offered very recently by Khan \cite{khan,khan2}. The author asserts that, in analogy with the mutually-enhanced-magicity (MEM) effect observed in predictions of masses with different energy-density functionals \cite{mem1,mem2}, the ISGMR energy in the doubly-magic nuclei might be anomalously higher. The obvious implication is that the calculations using interactions that are successful in describing the ISGMR in the doubly-magic nucleus $^{208}$Pb would necessarily overestimate the ISGMR energies in the open-shell nuclei. If this effect is manifested in any significant way, the energy of the ISGMR in the non-doubly-magic Pb isotopes, $^{204}$Pb and $^{206}$Pb, would be measurably lower than that in $^{208}$Pb~\cite{khan2}. In the only measurement of the ISGMR in $^{206}$Pb reported so far~\cite{mnh4}, this conjecture does not appear to hold. Still, precise measurements of the ISGMR in the Pb isotopes, using background-free inelastic spectra with high-energy $\alpha$ beams, would be worthwhile to fully examine this possibility. The incompressibility of a nucleus, $K_{A}$, may be expressed as: \begin{equation} K_{A} \sim K_{vol}(1 + cA^{-1/3}) + K_{\tau}((N - Z)/A)^{2} + K_{Coul}Z^{2}A^{-4/3} \end{equation} \noindent Here, c $\approx$ -1\cite{skp02}, and $K_{Coul}$ is essentially model independent (in the sense that the deviations from one theoretical model to another are quite small), so that the associated term can be calculated for a given isotope. Thus, for a series of isotopes, the difference $K_A-K_{Coul}Z^2A^{-4/3}$ may be approximated to have a quadratic relationship with the asymmetry parameter, of the type y = A + Bx$^2$, with K$_\tau$ being the coefficient, B, of the quadratic term. It has been established previously~\cite{shlomo2,pears} that direct fits to the Eq.~8 do not provide good constraints on the value of K$_\infty$. However, this expression is being used here not to obtain a value for K$_\infty$, but, rather, only to demonstrate the approximately quadratic relationship between K$_A$ and the asymmetry parameter. From such an analysis of the ISGMR data in the Sn isotopes, we have obtained a value of $K_\tau = -550\pm100$~MeV (see Fig. 4 in Ref.~\cite{tao}). This number is consistent with the value of $K_{\tau}=-370\pm120$~MeV obtained from an analysis of the isotopic transport ratios in medium-energy heavy-ion reactions \cite{bao3}. Incidentally, this value has been modified from the value of $-500\pm50$~MeV that was quoted previously by this group~\cite{exp-BAL-1,exp-BAL-4} and referred to in Ref.~\cite{tao}. It transpires that they had identified the quantity that they had obtained, $K_{asy}$, as being identical to $K_{\tau}$, the quantity that has been obtained from the ISGMR measurements; the two differ by a higher-order term \cite{sagawa6,bao3}. More recently, a value of $K_{\tau}=-500^{+125}_{-100}$~MeV has been obtained by Centelles {\em et al.} \cite{spain} from constraints put by neutron-skin data from anti-protonic atoms across the mass table; here again, it would appear that what the authors have termed $K_{\tau}$ is actually the aforementioned $K_{asy}$. Further, a value of $K_{\tau}=-500\pm50$ MeV has been obtained also by Sagawa {\em et al.} by comparing our Sn ISGMR data with calculations using different Skyrme Hamiltonians and RMF Lagrangians \cite{sagawa7}. The $K_{\tau}$ value obtained from our ISGMR measurements has, thus, been verified by a number of different procedures involving quite different data. A more precise determination of K$_\tau$ will likely result from extending the ISGMR measurements to longer isotopic chains. This provides strong motivation for measuring the ISGMR strength in unstable nuclei, a focus of current investigations at the new rare isotope beam facilities at RIKEN, GANIL, GSI, and NSCL~\cite{hb07,cm07,gsi,nscl}. Combined with the value of $K_\infty = 240\pm10$~MeV obtained from the ISGMR and ISGDR data~\cite{exp-ND-3,th-GC-3,th-JP-3,th-SS-2}, we now have ``experimental'' values of both K$_\infty$ and K$_\tau$ which, together, can provide a means of selecting the most appropriate of the interactions used in EOS calculations. For example, this combination of ``experimental'' values for K$_\infty$ and K$_\tau$ essentially rules out a vast majority of the Skyrme-type interactions currently in use in nuclear structure calculations~\cite{sagawa6,ugbulk}. Similar conclusions were reached for EOS equations in Refs.~\cite{exp-BAL-3,cs}. \section{Summary} We have measured the strength distributions of the isoscalar giant resonances (ISGMR, ISGDR, ISGQR, and HEOR) in the even-A $^{112-124}$Sn isotopes via inelastic scattering of 386-MeV $\alpha$ particles at extremely forward angles, including 0$^{\circ}$. The extracted parameters for these resonances are in good agreement with previously-obtained values where available. The ISGMR centroid energies are significantly lower than those predicted for these isotopes by recent calculations and point to the need for further theoretical exploration of applicable nuclear structure effects, especially the role of pairing in ISGMR strength calculations in the open-shell nuclei. The asymmetry-term, $K_{\tau}$, in the expression for the nuclear incompressibility has been determined to be $-550 \pm 100$ MeV from the ISGMR data in Sn isotopes and is found to be consistent with a number of indirectly extracted values for this parameter. \section{Acknowledgments} We wish to thank the RCNP staff for providing high-quality $\alpha$ beams required for these measurements. This work has been supported in part by the US-Japan Cooperative Science Program of the JSPS, and by the National Science Foundation (Grants No. INT03-42942, PHY04-57120, and PHY07-58100).
\section{Introduction}\label{s:intro} Very massive stars appear to go through a phase of instability and large mass loss; during this stage, a star is a member of the luminous blue variable (LBV) class (see \citealt{Humphreys94} for a review). In addition to low-amplitude variability (called S~Dor variability after the prototypical LBV), where the star ejects mass from its envelope but its bolometric luminosity remains nearly constant, some LBVs have ``giant eruptions.'' Giant eruptions can expel $\gtrsim 1 M_{\sun}$ of material, while having a luminosity similar to the lowest-luminosity supernovae (SNe). The classical examples of giant eruptions are P~Cygni in 1600 and $\eta$~Car in 1843, and more recent, extragalactic examples include SN~1961V (\citealt{Goodrich89, Filippenko95, VanDyk02}; but see \citealt{Chu04}), V12/SN~1954J \citep{Smith01, VanDyk05}, SN~1997bs \citep{VanDyk00}, SN~2000ch \citep{Wagner04}, and V37/SN~2002kg \citep{Weis05, Maund06, VanDyk06}. Other potential examples exist, but all events listed above have pre-event imaging where the progenitor star has been identified as a probable or definite LBV. Two examples of a new class of stellar eruptions have recently emerged. The progenitors of NGC~300 OT2008-1\ and SN~2008S were both detected in pre-event {\it Spitzer} images, but were undetected to deep limits in the optical \citep{Prieto08, Berger09:ngc, Bond09}, indicating significant reddening from circumstellar dust. The progenitor stars were originally believed to have ZAMS masses of 8 -- 20~$M_{\sun}$, which is below that of the least massive LBVs. Using the stars in the vicinity of NGC~300 OT2008-1, \citet{Gogarten09} found a slightly higher mass range of 12 -- 25~$M_{\sun}$. A reasonable range for the initial mass of the progenitors is $\sim\!\!$~ 10 -- $25 M_{\sun}$, with NGC~300 OT2008-1, having a more luminous progenitor, being toward the upper end of that range. Recently, two transients were discovered with one being very similar to classical LBV giant eruptions (SN~2009ip), and another sharing characteristics of both LBV giant eruptions and the outbursts of the dusty stars discussed above (UGC~2773 OT2009-1). SN~2009ip\ was discovered by \citet{Maza09} in NGC~7259 ($\mu = 32.05 \pm 0.15$~mag\footnote{We use the distance modulus corresponding to the Hubble distance for NGC~7259 with $H_{0} = 73$~km~s$^{-1}$~Mpc$^{-1}$ and correcting for the Virgo Infall and Great Attractor flow model of \citet{Mould00}. \citet{Smith09:outburst} use a distance modulus of 31.55~mag, which differs by 0.50~mag from our assumed value, corresponding to the Hubble-flow distance modulus correcting only for the CMB dipole.}; $D \approx 24$~Mpc) on 2009 August 26 (UT dates will be used throughout this paper). \citet{Miller09} noted that SN~2009ip\ had been variable for several years and identified a possible progenitor with $M_{\rm F606W} \approx -10.1$~mag, in an archival {\it Hubble Space Telescope} (\protect\hbox{\it HST} ) image. The variability and high luminosity of the event led \citet{Miller09} to suggest that SN~2009ip\ was either an LBV or cataclysmic variable outburst. An early spectrum of the event showed a blue continuum with relatively narrow (${\rm FWHM} = 550$~km~s$^{-1}$ ) H Balmer lines. The combination of the spectrum with the relatively low absolute magnitude ($R \approx -13.7$~mag) led us to conclude that SN~2009ip\ was an LBV giant eruption \citep*{Berger09:ip}. The transient underwent extreme variability shortly after maximum\footnote{To be consistent with \citetalias{Smith09:outburst}, we adopt ${\rm MJD} = 55061.5$ and ${\rm MJD} = 55071.75$ as times of maximum light for UGC~2773 OT2009-1\ and SN~2009ip, respectively. However, we note that the objects may reach their true maximum later, which UGC~2773 OT2009-1\ already has (as shown by data presented in \citetalias{Smith09:outburst})}, fading by at least 3~mag in 16~days and rebrightening by 2~mag in the next 10~days \citep{Li09}, reminiscent of the variability immediately before maximum of the 1843 eruption of $\eta$~Car \citep{Frew04} and immediately after maximum in the LBV outburst SN~2000ch \citep{Wagner04}. \citet[][hereafter S09]{Smith09:outburst} presented a historical light curve of SN~2009ip\ that begins 5~years before maximum light, excluding the \protect\hbox{\it HST} \ image of the progenitor. The star varied by at least 1.5~mag during this time. \citetalias{Smith09:outburst} presented additional data which led them to conclude that SN~2009ip\ was the giant eruption of a 50 -- $80 M_{\sun}$ LBV. UGC~2773 OT2009-1\ was discovered by \citet{Boles09} in UGC~2773 ($\mu = 28.82 \pm 0.17$~mag; $D \approx 6$~Mpc) on 2009 August 18. It was originally reported as a possible SN. We obtained a spectrum and noted that it had a peculiar spectrum with relatively narrow (${\rm FWHM} = 350$~km~s$^{-1}$ ) H$\alpha$ emission, P-Cygni lines from the \ion{Ca}{2} NIR triplet, and [\ion{Ca}{2}] emission lines \citep{Berger09:ugc}. We also noted that the spectrum was similar to that of NGC~300 OT2008-1\ and mentioned that it was possibly a very low-luminosity SN~II or an LBV outburst, ``but the strong [\ion{Ca}{2}] emission would be unexpected in this case.'' \citet{Berger09:ugc} also detected a potential progenitor star in archival \protect\hbox{\it HST} \ images. \citetalias{Smith09:outburst} presented a historical light curve of SN~2009ip\ that begins 9~years (excluding the \protect\hbox{\it HST} \ image of the progenitor) before maximum light. The star slowly increased from ${\rm F814W} = 22.22$~mag to an unfiltered magnitude of 17.70~mag at maximum light, corresponding to a linear increase of $\sim\!\!$~ 0.4~mag~year$^{-1}$ before outburst. \citetalias{Smith09:outburst} concluded from the progenitor identification, their historic light curve, the peak luminosity, and optical light curves that UGC~2773 OT2009-1\ was the outburst of a $\gtrsim 20 M_{\sun}$ LBV. This is the first in a series of papers where we investigate the diversity of massive star outbursts. In this paper we demonstrate the heterogeneity of the class with observations of SN~2009ip\ and UGC~2773 OT2009-1. In future papers, we will detail the properties of the class and the links between observations and the physical mechanisms which cause the outbursts. In Section~\ref{s:obs}, we present ultraviolet (UV), optical, and near-infrared (NIR) photometry and optical and NIR spectroscopy of UGC~2773 OT2009-1\ and SN~2009ip. In this section, we also refine previous identifications of the progenitors. In Section~\ref{s:results}, we examine the progenitor masses, the spectroscopic characteristics of the outbursts, and the spectral-energy distributions (SEDs) of the events. In Section~\ref{s:disc}, we discuss how these outbursts connect to previous massive star outbursts and the mass loss history and ultimate fates of massive stars. We summarize our conclusions in Section~\ref{s:con}. \section{Observations and Data Reduction}\label{s:obs} \subsection{Identification and \protect\hbox{\it HST} \ Photometry of the Progenitors} UGC~2773 was observed with \protect\hbox{\it HST} \!\!/WFPC2 on 1999 August 14 (Program 8192; PI Seitzer). The observations included two exposures of 600~s each with the F606W and F814W (roughly $V$ and $I$) filters. NGC~7259 was observed with \protect\hbox{\it HST} \!\!/WFPC2 on 1999 June 29 (Program 6359; PI Stiavelli). Exposures of 200 and 400~s were obtained with the F606W filter. To determine whether the progenitors of SN~2009ip\ and UGC~2773 OT2009-1\ are detected in the archival \protect\hbox{\it HST} \ observations, we performed differential astrometry using optical observations of the transients. Observations of UGC~2773 OT2009-1\ were obtained with the Gemini Multi-Object Spectrograph (GMOS) on the Gemini-North 8-m telescope, and the astrometry was performed using 55 objects in common with the \protect\hbox{\it HST} \!\!/WFPC2 images resulting in an astrometric rms of $\sigma_{\rm GB\rightarrow HST} = 24$~mas in each coordinate. Observations of SN~2009ip\ were obtained with the Inamori Magellan Areal Camera and Spectrograph (IMACS) on the Magellan/Baade 6.5-m telescope, and the astrometry was performed using 10 objects in common with the \protect\hbox{\it HST} \!\!/WFPC2 image resulting in an astrometric rms of $\sigma_{\rm GB\rightarrow HST} = 38$~mas in each coordinate. The positions of the two transients on the archival \protect\hbox{\it HST} \ images are shown in Figure~\ref{f:proj}. In both cases, we find a clear coincidence with objects in the archival \protect\hbox{\it HST} \ images. For SN~2009ip\ we find an offset of $24 \pm 38$~mas relative to the object in the WFPC2/F606W image, while for UGC~2773 OT2009-1\ we find an offset of $32 \pm 24$~mas relative to the object in the WFPC2/F606W and F814W images. \begin{figure*} \begin{center} \epsscale{1.15} \plottwo{paper_ugc2773.eps}{paper_sn2009ip.eps} \caption{{\it HST}/WFPC2 F606W image at the position of UGC~2773 OT2009-1\ (left) and SN~2009ip\ (right) obtained 10~years before maximum. Both images are $10\hbox{$^{\prime\prime}$} \times 10\hbox{$^{\prime\prime}$}$, and North is up and East is left. The UGC~2773 OT2009-1\ and SN~2009ip\ images have pixel scales of 0.1\hbox{$^{\prime\prime}$}~pixel$^{-1}$. The position of each transient is marked by the black circle whose radius corresponds to $10 \sigma$ uncertainty in the position.}\label{f:proj} \end{center} \end{figure*} The measurements of the photometry for UGC~2773 OT2009-1\ and nearby stars were done using HSTphot 1.1 \citep{Dolphin00}. HSTphot was run using a weighted PSF-fit, which is recommended for crowded fields, and a local sky determination, which is recommended for rapidly varying backgrounds. HSTphot performs the conversion from \protect\hbox{\it HST} \!\!/WFPC2 flight magnitudes to the Bessel magnitude system. Our astrometry and photometry for the nominal progenitor of UGC~2773 OT2009-1\ and nearby stars are listed in Table~\ref{t:hst}. \begin{deluxetable*}{lllll} \tabletypesize{\scriptsize} \tablewidth{0pt} \tablecaption{{\it HST} Photometry of Stars Near UGC~2773 OT2009-1\label{t:hst}} \tablehead{ \colhead{Object} & \colhead{R.A.} & \colhead{Dec.} & \colhead{F606W (mag)} & \colhead{F814W (mag)}} \startdata \phn1\tablenotemark{a} & 03:32:7.240 & +47:47:39.60 & 22.824 (0.032) & 22.286 (0.053) \\ \phn2 & 03:32:7.258 & +47:47:39.99 & 22.474 (0.027) & 22.164 (0.050) \\ \phn3 & 03:32:7.229 & +47:47:40.14 & 22.672 (0.036) & 22.222 (0.054) \\ \phn4 & 03:32:7.203 & +47:47:40.29 & 22.853 (0.035) & 22.591 (0.081) \\ \phn5 & 03:32:7.162 & +47:47:40.23 & 23.864 (0.072) & 22.729 (0.074) \\ \phn6 & 03:32:7.132 & +47:47:39.94 & 23.413 (0.051) & 22.748 (0.092) \\ \phn7 & 03:32:7.048 & +47:47:39.45 & 22.925 (0.034) & 20.956 (0.020) \\ \phn8 & 03:32:7.225 & +47:47:40.79 & 21.724 (0.016) & 21.043 (0.022) \\ \phn9 & 03:32:7.266 & +47:47:40.51 & 23.007 (0.039) & 22.548 (0.067) \\ 10 & 03:32:7.377 & +47:47:40.44 & 22.967 (0.035) & 21.682 (0.035) \\ 11 & 03:32:7.230 & +47:47:38.60 & 23.762 (0.092) & 23.182 (0.113) \\ 12 & 03:32:7.263 & +47:47:38.34 & 23.578 (0.058) & 23.095 (0.100) \\ 13 & 03:32:7.306 & +47:47:38.61 & 24.015 (0.081) & 23.304 (0.116) \\ 14 & 03:32:7.409 & +47:47:39.44 & 21.194 (0.026) & 19.934 (0.017) \enddata \tablenotetext{a}{Star~1 is identified as the progenitor of UGC~2773 OT2009-1.} \end{deluxetable*} We performed photometry of the point source coincident with SN~2009ip\ using a $0.5\hbox{$^{\prime\prime}$}$ aperture and a zero-point of 22.47~mag appropriate for the F606W filter. We further applied a correction of $-0.29$~mag to convert to the Vega system, and applied a correction for the Galactic extinction of 0.05~mag. The resulting magnitude of the source is $21.84\pm 0.25$~mag. For reasonable colors (see Section~\ref{sss:ip_mass}), this corresponds to $M_{V} = -10.3$~mag. \subsection{Ultraviolet, Optical, and Near-Infrared Photometry} We obtained optical photometry of UGC~2773 OT2009-1\ with the Gemini Multi-Object Spectrograph (GMOS) on the Gemini-North 8-m telescope in the \protect\hbox{$gri$} \ filters. We performed the photometry using IRAF/{\tt phot} with the standard GMOS zero-points\footnote{\tt http://www.gemini.edu/sciops/instruments/gmos/imaging}. Our results are presented in Table~\ref{t:phot}. \begin{deluxetable*}{rlcll} \tabletypesize{\scriptsize} \tablewidth{0pt} \tablecaption{UV and Optical Photometry of UGC~2773 OT2009-1\ and SN~2009ip\label{t:phot}} \tablehead{ \colhead{Object} & \colhead{MJD} & \colhead{Filter} & \colhead{Mag} & \colhead{Telescope}} \startdata UGC~2773 OT2009-1 & 51404.13 & F606W & 22.82 (0.03) & {\it HST} \\ UGC~2773 OT2009-1 & 51404.14 & F814W & 22.29 (0.05) & {\it HST} \\ UGC~2773 OT2009-1 & 55078.38 & $J$ & 15.47 (0.06) & Fan Mountain \\ UGC~2773 OT2009-1 & 55078.40 & $H$ & 14.99 (0.06) & Fan Mountain \\ UGC~2773 OT2009-1 & 55078.39 & $K_{s}$ & 14.63 (0.07) & Fan Mountain \\ UGC~2773 OT2009-1 & 55078.53 & $g$ & 18.32 (0.01) & Gemini-North \\ UGC~2773 OT2009-1 & 55078.53 & $r$ & 17.22 (0.01) & Gemini-North \\ UGC~2773 OT2009-1 & 55078.53 & $i$ & 16.68 (0.01) & Gemini-North \\ UGC~2773 OT2009-1 & 55089.31 & $J$ & 15.47 (0.06) & Fan Mountain \\ UGC~2773 OT2009-1 & 55089.32 & $H$ & 14.91 (0.07) & Fan Mountain \\ UGC~2773 OT2009-1 & 55089.33 & $K_{s}$ & 14.91 (0.09) & Fan Mountain \\ \tableline SN~2009ip & 51358.50 & F606W & 21.84 (0.17) & {\it HST} \\ SN~2009ip & 55084.44 & UVW2 & 21.09 (0.19) & {\it Swift} \\ SN~2009ip & 55084.45 & UVM2 & 20.92 (0.28) & {\it Swift} \\ SN~2009ip & 55084.44 & UVW1 & 20.69 (0.18) & {\it Swift} \\ SN~2009ip & 55084.44 & $U$ & 20.29 (0.16) & {\it Swift} \\ SN~2009ip & 55084.44 & $B$ & 20.64 (0.09) & {\it Swift} \\ SN~2009ip & 55084.45 & $V$ & 20.47 (0.37) & {\it Swift} \enddata \end{deluxetable*} We obtained near-infrared (NIR) photometry of UGC~2773 OT2009-1\ with FanCam, a $1024 \times 1024$ HAWAII-I HgCdTe imaging system on the University of Virginia's 31-inch telescope at Fan Mountain, just outside of Charlottesville, VA \citep{Kanneganti09}. Each epoch consists of fifteen minutes of integration in \protect\hbox{$J\!H\!K_{s}$} \ bands, which have detection limits at the $10 \sigma$ level of 0.066, 0.098, and 0.156~mJy (or 18.5, 17.5, and 16.5~mag), respectively. Individual exposures are sky-background limited and have an integration time of either 30 or 60~s. Flat-field frames are composed of dusk and dawn sky observations. We employed standard NIR data reduction techniques in IRAF\footnote{IRAF is distributed by the National Optical Astronomy Observatory, which is operated by the Association of Universities for Research in Astronomy (AURA) under cooperative agreement with the National Science Foundation.}. Because of the relatively small galaxy size, it was possible to fit the entire galaxy in a single array quadrant. Empty quadrants were efficiently utilized as sky exposures. Data were taken with the galaxy placed in each quadrant and each quadrant was reduced separately. Ultimately, all reduced quadrants were coadded. We performed photometry with IRAF's PSF package. For magnitude calibration, the transient is compared to 2MASS reference stars located in the field of view. Table~\ref{t:phot} lists our \protect\hbox{$J\!H\!K_{s}$} \ photometry, which is similar to the single epoch \protect\hbox{$J\!H\!K_{s}$} \ data from \citetalias{Smith09:outburst}. We obtained UV and optical observations of SN~2009ip\ with the {\it Swift} UV/optical telescope on 2009 September 10. The data were processed using standard routines within the HEASOFT package. Photometry of the transient in all filters, with the exception of UVW2, was performed using a 2\hbox{$^{\prime\prime}$}\ aperture to avoid contamination from nearby objects. Aperture corrections to the standard 5\hbox{$^{\prime\prime}$}\ aperture were determined using isolated stars; photometry of the source in the UVW2 filter was performed using a 5\hbox{$^{\prime\prime}$}\ aperture. \subsection{X-ray Observations} We observed SN~2009ip\ and UGC~2773 OT2009-1\ with the {\it Swift} X-ray Telescope on 2009 September 10 for a total exposure time of 9.0 and 4.2~ks, respectively. No X-ray counterpart is detected at the position of either source to a limit of $F_{X} \lesssim 2.8 \times 10^{-14}$ and $\lesssim 1.1 \times 10^{-13}$~erg~s$^{-1}$~cm$^{-2}$, respectively ($95\%$ limit). In both cases we assume a power law model with an electron index of $-2$, and account for the Galactic neutral hydrogen column. The corresponding limits on the luminosity are $L_{X} \lesssim 1.9 \times 10^{39}$ and $\lesssim 4.8 \times 10^{38}$~erg~s$^{-1}$. These limits are comparable to the X-ray emission from SNe on a similar timescale \citep[e.g.,][]{Soderberg08}. \subsection{Radio Observations} We observed the both LBV candidates with the Very Large Array\footnote{The Very Large Array is operated by the National Radio Astronomy Observatory, a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc.} (VLA) following their optical discovery to search for radio counterparts, under Rapid Response programs AS1001 and AS1002 (PI Soderberg). Our radio observations were carried out at two frequencies, 8.46 and 22.5 GHz, on dates spanning 2009 September 7.36 - 16.51 in the C-array antenna configuration. All observations were taken in standard continuum observing mode with a bandwidth of $2 \times 50$~MHz. Phase referencing was performed with calibrators J0325+469 and J2213-254, and we used 3C38 (J0137+331) for flux calibration. Data were reduced using standard packages within the Astronomical Image Processing System (AIPS). We detect no radio sources in positional coincidence with either object and derive upper limits summarized in Table~\ref{t:vla}. At 8.5 GHz, our upper limits\footnote{Upper limits are calculated as the measured flux density at the optical position summed with $2 \times$ {\it rms} the off-source map noise.} correspond to $L_{\nu} < 1.3 \times 10^{26}$~erg~s$^{-1}$~Hz$^{-1}$ and $L_{\nu} < 2.6 \times 10^{24}$~erg~s$^{-1}$~Hz$^{-1}$ for SN~2009ip\ and UGC~2773 OT2009-1, respectively. These limits are less luminous than an extrapolation of the observed SN~1961V radio emission at $t \approx 10$~years, to a similarly early epoch as $L_{\nu} \propto t^{-1.75}$ \citep{Stockdale01}. We note, however, that the SN~1961V radio emission may have reached maximum intensity significantly later than our observations of UGC~2773 OT2009-1\ and SN~2009ip, similar to the radio evolution of SNe~IIn which typically reach maximum light several years after the explosion \citep{VanDyk96}. \begin{deluxetable}{llrr} \tablecaption{VLA observations of UGC~2773 OT2009-1\ and SN~2009ip\label{t:vla}} \tablewidth{0pt} \tablehead{ \colhead{Object} & \colhead{Date} & \colhead{$F_{\nu, 8.5}$} & \colhead{$F_{\nu, 22}$} \\ \colhead {} & \colhead{(UT)} & \colhead{($\mu$Jy)} & \colhead{($\mu$Jy)} } \startdata SN~2009ip & 2009 Sep 7.36 & \nodata & $-122 \pm 243$ \\ \nodata & 2009 Sep 9.24 & $67 \pm 47$ & \nodata \\ UGC~2773 OT2009-1 & 2009 Sep 13.57 & \nodata & $-57 \pm 66$ \\ \nodata & 2009 Sep 16.51 & $11 \pm 27$ & \nodata \enddata \tablecomments{Uncertainties are $1\sigma$ {\it rms} map noise.} \end{deluxetable} A comparison of these radio upper limits for the outbursts to the observed properties of other core-collapse SNe places them among the least luminous events, 2-4 orders of magnitude less luminous than the most powerful SNe~IIn, and 4-200 times higher than the early radio signal seen for SN~1987A \citep{Ball95}. Through this simple comparison we emphasize that radio data alone cannot distinguish between massive star outbursts and catastrophic explosions. However, with the 10-fold increase in continuum sensitivity provided by the EVLA we will begin to map out the radio properties for massive star outbursts and enable direct comparisons with those of core-collapse SNe (e.g., NRAO Key Project AS1020, "Exotic Explosions,Eruptions,and Disruptions: A New Transient Phase-Space", PI Soderberg). \subsection{Optical Spectroscopy}\label{ss:opt_spec} We obtained low- and medium-resolution spectra of SN~2009ip\ and UGC~2773 OT2009-1\ with the MagE spectrograph \citep{Marshall08} on the Magellan Clay 6.5~m telescope, the Blue Channel spectrograph \citep*{Schmidt89} on the MMT 6.5~m telescope, and GMOS \citep{Hook04} on the Gemini-North 8~m telescope. A journal of our optical spectroscopic observations can be found in Table~\ref{t:spec}. \begin{deluxetable*}{lllcrl} \tabletypesize{\scriptsize} \tablewidth{0pt} \tablecaption{Log of Optical Spectral Observations\label{t:spec}} \tablehead{ \colhead{} & \colhead{} & \colhead{Telescope /} & \colhead{Grating /} & \colhead{Exposure} & \colhead{} \\ \colhead{Phase\tablenotemark{a}} & \colhead{UT Date} & \colhead{Instrument} & \colhead{Central Wavelength (\AA)} & \colhead{(s)} & \colhead{Observer\tablenotemark{b}}} \startdata \cutinhead{UGC~2773 OT2009-1} 15.1 & 2009 Sep.\ 2.6 & Gemini/GMOS & R400/7000 & $2 \times 1200$ & KO, RM \\ 32.9 & 2009 Sep.\ 20.4 & MMT/Blue Channel & 300/5787 & $3 \times 1200$ & PC \\ 33.9 & 2009 Sep.\ 21.4 & MMT/Blue Channel & 300/5787 & 1200 & PC \\ 33.9 & 2009 Sep.\ 21.4 & MMT/Blue Channel & 832/4029 & $2 \times 900$ & PC \\ 34.0 & 2009 Sep.\ 21.5 & MMT/Blue Channel & 832/4830 & $2 \times 900$ & PC \\ 34.0 & 2009 Sep.\ 21.5 & MMT/Blue Channel & 832/6563 & $3 \times 900$ & PC \\ 95.8 & 2009 Nov.\ 22.3 & MMT/Blue Channel & 300/5787 & $2 \times 1800$ & PC \\ \cutinhead{SN~2009ip} 3.5 & 2009 Sep.\ 1.3 & Clay/MagE & \nodata & $3 \times 900$ & II \\ 22.5 & 2009 Sep.\ 20.3 & MMT/Blue Channel & 300/5787 & $4 \times 1200$ & PC \\ 23.5 & 2009 Sep.\ 21.2 & MMT/Blue Channel & 832/4029 & $2 \times 1200$ & PC \\ 23.5 & 2009 Sep.\ 21.3 & MMT/Blue Channel & 832/4830 & $2 \times 1200$ & PC \\ 23.5 & 2009 Sep.\ 21.3 & MMT/Blue Channel & 832/6563 & 1200 & PC \\ 85.6 & 2009 Nov.\ 22.1 & MMT/Blue Channel & 300/5787 & $3 \times 1200$ & PC \enddata \tablenotetext{a}{Days since maximum, MJD 55,061.5 and 55,071.8 for UGC~2773 OT2009-1\ and SN~2009ip\ \citepalias{Smith09:outburst}, respectively.} \tablenotetext{b}{II = I.\ Ivans, KO = K.\ Olsen, PC = P.\ Challis, RM = R.\ McDermid} \end{deluxetable*} Standard CCD processing and spectrum extraction were accomplished with IRAF. The data were extracted using the optimal algorithm of \citet{Horne86}. Low-order polynomial fits to calibration-lamp spectra were used to establish the wavelength scale, and small adjustments derived from night-sky lines in the object frames were applied. For the MagE spectra, the sky was subtracted from the images using the method described by \citet{Kelson03}. The GMOS data were reduced using the Gemini IRAF package \citep[for details, see ][]{Foley06}. We employed our own IDL routines to flux calibrate the data and remove telluric lines using the well-exposed continua of the spectrophotometric standards \citep{Wade88, Foley03, Foley09:08ha}. Representative spectra of SN~2009ip\ and UGC~2773 OT2009-1\ are presented in Figure~\ref{f:opt_spec}. Both objects have similar blue continua, but the line features are very different. SN~2009ip\ has few line features besides strong H Balmer lines, Na~D, and \ion{He}{1}. Although UGC~2773 OT2009-1\ has a strong H$\alpha$ line, it is much narrower and weaker than that of SN~2009ip. UGC~2773 OT2009-1\ also displays many additional narrow line features, including lines from intermediate mass and Fe-group elements. Blueward of $\sim\!\!$~ 5500~\AA, UGC~2773 OT2009-1\ is dominated by a forest of \ion{Fe}{2} lines. Finally, UGC~2773 OT2009-1\ has very strong \ion{Ca}{2} NIR triplet lines and [\ion{Ca}{2}] $\lambda\lambda 7291$, 7324 lines. These features, and particularly [\ion{Ca}{2}] are rarely seen in classical LBV outbursts, but were distinguishing features of SN~1999bw \citep{Garnavich99}, SN~2008S \citep[e.g.,][]{Smith09:08s}, and NGC~300 OT2008-1\ \citep[e.g.,][]{Berger09:ngc}. \begin{figure*} \begin{center} \epsscale{0.6} \rotatebox{90}{ \plotone{lbv_opt.ps}} \caption{Optical spectra of SN~2009ip\ and UGC~2773 OT2009-1. The spectrum of UGC~2773 OT2009-1\ has been dereddened by $E(B-V) = 0.564$~mag. Several lines are identified and marked.}\label{f:opt_spec}. \end{center} \end{figure*} The spectra from 2009 September 21 (corresponding to days 34 and 24 for UGC~2773 OT2009-1\ and SN~2009ip, respectively) were obtained on the same night as the LRIS spectra shown by \citetalias{Smith09:outburst}. \subsection{Near-Infrared Spectroscopy} On 2009 September 9 (22~days after maximum), we obtained a 2400~s NIR spectrum of UGC~2773 OT2009-1\ with TripleSpec, a medium resolution NIR spectrograph located at Apache Point Observatory. This spectrograph is one of three NIR, cross-dispersed spectrographs covering wavelengths from 1 -- 2.4~\micron\ simultaneously at a resolution of $\sim\!\!$~ 3500 \citep{Wilson04, Herter08}. We collected eight, 300-s sky-background limited exposures, for a total integration time of 2400~s. We extracted the spectrum with a modified version of the IDL-based SpexTool \citep*{Cushing04}. This tool removes any contribution from the underlying galactic arm by fitting the background with a 2nd order polynomial. \section{Results}\label{s:results} \subsection{Progenitor Masses} \subsubsection{SN~2009ip}\label{sss:ip_mass} We use the absolute magnitude, $M_{V} = -10.3$~mag, of the progenitor of SN~2009ip, along with an estimated range of $V - I$ colors of $-0.05$ to $1.4$~mag (representative of LBV colors spanning from O to F spectral types), to plot it as a line on a color-magnitude diagram (Figure~\ref{f:cmd}). For this progenitor we find a Milky Way extinction of $E(B-V) = 0.019$~mag ($A_{V} = 0.05$ mag; \citealt*{Schlegel98}). We adopt a distance modulus of $\mu = 32.05$~mag for NGC~7259 (see discussion in Section~\ref{s:intro}), and assume no additional host galaxy or circumstellar extinction. \begin{figure} \begin{center} \epsscale{1.23} \plotone{HRD_final.ps} \caption{Color-magnitude diagram ($V-I$ vs.\ $M_{V}$) for the progenitor of UGC~2773 OT2009-1\ (star) and stars spatially located within the same star cluster (grey circles). The measurements have been corrected for the Milky Way extinction of $A_{V} = 1.75$~mag and $E(V-I) = 0.902$~mag, but no host or circumstellar extinction is assumed for the stars. For comparison, solar metallicity, non-rotating, ``standard'' mass loss stellar evolution tracks are also plotted \citep{Schaller92}. The progenitor of UGC~2773 OT2009-1\ has the same colors and absolute magnitude of a $20 M_{\sun}$ model. The stars in the cluster are consistent with the models of stars with ZAMS masses $\leq 25 M_{\sun}$, but a single star is also consistent with a much higher mass. The arrow represents $A_{V} = 0.5$~mag of additional extinction (assuming $R_{V} = 3.1$). The dashed line represents the absolute magnitude of the progenitor of SN~2009ip\ with a reasonable range of colors. The luminosity of the progenitor of SN~2009ip\ is consistent with an initial mass of $\gtrsim 60 M_{\sun}$.}\label{f:cmd} \end{center} \end{figure} In Figure~\ref{f:cmd}, we compare the color of the SN~2009ip\ progenitor to the non-rotating, standard mass-loss evolutionary tracks of the Geneva group \citep{Schaller92}. From this plot we can place a lower initial mass limit of $60 M_{\sun}$ on the progenitor of SN~2009ip\ in the absence of a color estimate for this progenitor, the higher-mass evolutionary tracks all coincide with its estimated location on the color-magnitude diagram, precluding us from placing an upper limit on this initial mass estimate. Figure~\ref{f:cmd} assumes a solar metallicity for these tracks; however, we find that our progenitor mass prediction is consistent across the full range of metallicities accommodated by the Geneva evolutionary tracks ($Z = 0.05 Z_{\sun}$ to $Z = 2 Z_{\sun}$). It should be noted that an increased amount of extinction, from the host galaxy or circumstellar environment, could also effectively increase the estimated initial mass of this progenitor. \citetalias{Smith09:outburst} estimated the initial mass for the progenitor of SN~2009ip\ to be 50 -- $80 M_{\sun}$. Although the {\it HST} photometry from \citetalias{Smith09:outburst} is the same as that presented here, their assumed distance modulus is 0.50~mag smaller than our assumed value. They also make no color correction to transform the F606W measurements into $V$. Despite these differences, the two mass ranges are similar. \subsubsection{UGC~2773 OT2009-1} Using its $M_{V}$ and $V-I$ color, we are able to determine an estimate of the initial mass for the progenitor of UGC~2773 OT2009-1\ (Figure~\ref{f:cmd}). There is significant Milky Way extinction of $E(B-V) = 0.564$~mag ($A_{V} = 1.75$~mag; \citealt*{Schlegel98}), which we convert to $E(V-I) = 0.902$~mag \citep{Schultz75}. We use a distance modulus of $\mu = 28.82$~mag and initially assume no host galaxy or circumstellar extinction. From Figure~\ref{f:cmd}, we find that the progenitor of UGC~2773 OT2009-1\ is consistent with an initial mass of $\sim\!\!$~ $20 M_{\sun}$. Our progenitor mass prediction remains the same across the full range of metallicities covered by the Geneva evolutionary tracks, and is consistent with the value found by \citetalias{Smith09:outburst}. We have also performed this procedure on several stars in the vicinity of the progenitor of UGC~2773 OT2009-1. Assuming that all of these stars are part of a cluster and were formed at the same time, they should place additional limits on the current maximum-mass stars of the cluster. These stars are all consistent with an initial mass of $M_{\sun} \lesssim 25 M_{\sun}$. There is a single star that is particularly blue (and therefore potentially very massive), but it is still consistent with an initial mass of $25 M_{\sun}$. The likely association of the progenitor of UGC~2773 OT2009-1\ with this cluster and its upper mass limit of $\sim\!\!$~ 25~$M_{\sun}$ further supports the initial mass estimate for the progenitor of UGC~2773 OT2009-1. Considering the blue colors of the stars in the cluster, it is unlikely that they are significantly reddened by host galaxy dust. As shown in Figure~\ref{f:cmd}, a relatively small amount of extinction could significantly increase our initial mass estimate for UGC~2773 OT2009-1. In Section~\ref{ss:sed} we show that there was likely a significant amount of circumstellar dust existing before the outburst, indicating that the progenitor had an initial mass much larger than the reddening-free estimate of $20 M_{\sun}$. The combination of the reddening-free initial mass estimate for the progenitor of UGC~2773 OT2009-1, the initial mass estimates of stars likely within the same cluster as the progenitor, and the probably circumstellar dust extinction give us a conservative lower limit on the initial mass of the progenitor of UGC~2773 OT2009-1\ of $\sim\!\!$~ $25 M_{\sun}$. \subsection{Spectroscopic Comparisons} \subsubsection{SN~2009ip}\label{sss:ip_spec} We present the 24 and 86~day spectra of SN~2009ip\ in Figure~\ref{f:ip_spec}. In the upper panel of Figure~\ref{f:ip_spec}, the 24~day spectrum is compared to the 2~day spectrum of the LBV outburst SN~1997bs \citep{VanDyk00}. Both objects have blue continua, strong and narrow H Balmer lines, and \ion{He}{1} and \ion{Fe}{2} emission lines. Unlike SN~1997bs, SN~2009ip\ has a particularly strong \ion{He}{1} $\lambda 5876$ line (with some possible contribution from Na~D), and all H Balmer lines and \ion{He}{1} $\lambda 5876$ show strong absorption features with minima blueshifted by $\sim\!\!$~ 3000~km~s$^{-1}$ \ (see Section~\ref{sss:halpha} for a detailed discussion of this high-velocity absorption). \begin{figure} \begin{center} \epsscale{0.9} \rotatebox{90}{ \plotone{sn2009ip_comp.ps}} \caption{Top panel: the 23~day optical spectrum of SN~2009ip\ compared to the 2~day spectrum of SN~1997bs \citep{VanDyk00}. Bottom panel: the 86~day optical spectrum of SN~2009ip\ after smoothing and subtracting a 10,000~K blackbody (see text for details). For comparison, the 25~day spectrum of SN~1998S (after subtracting a 10,000~K blackbody) is also shown \citep{Leonard00}. Prominent, high-velocity lines have been marked.}\label{f:ip_spec} \end{center} \end{figure} Inspecting the 25~day spectrum from \citetalias{Smith09:outburst}, we see some indication of the 3000~km~s$^{-1}$ \ absorption component, particularly for H$\beta$ and \ion{He}{1} $\lambda 5876$; however, the absorption is much stronger in the spectra presented here (which were obtained at very similar times). We have reduced our spectra with many different extraction regions and backgrounds, with the high-velocity absorption features are present in all reductions. The same feature is present in all spectra taken with the MMT on days 23 and 24, which were taken with different gratings and wavelength regions. It is also present on day 86, but with a different velocity. The absorption is present for all Balmer lines and \ion{He}{1} $\lambda 5876$. Despite the apparent differences with the concurrent Keck spectrum \citepalias{Smith09:outburst}, we are confident that the high-velocity absorption features are real and not an artifact of the spectral reductions. The bottom panel of Figure~\ref{f:ip_spec} displays the spectra of SN~2009ip\ and SN~IIn~1998S (after subtracting a 10,000~K blackbody spectrum from both) from 86 and 25~days, respectively. An inverse-variance weighted Gaussian filter (with a width of 1000~km~s$^{-1}$ ) has been applied to the spectrum of SN~2009ip\ \citep{Blondin06}. This filtering will smear out features with intrinsic widths less than 1000~km~s$^{-1}$ , but will appropriately smooth features on larger scales. The high-velocity absorption in the 86~day spectrum of SN~2009ip\ is at a {\it higher} velocity than at 24~days. At this epoch, the velocity of the fast-moving SN~2009ip\ ejecta are very similar to that of SN~1998S. Although the H Balmer emission lines are much stronger in SN~2009ip, most other features are similar in the two spectra. In particular, SN~2009ip\ shows the H Balmer, \ion{He}{1}, \ion{Sc}{2}, and \ion{Fe}{2} features seen in SN~1998S. SN~2009ip\ is missing the strong absorption at 6250~\AA\ that is attributed to \ion{Si}{2} in SN~1998S \citep{Leonard00}. This feature may be the result of a significant amount of nuclear burning, and thus not present in the ejecta of SN~2009ip. \subsubsection{UGC~2773 OT2009-1} As discussed in Section~\ref{ss:opt_spec}, UGC~2773 OT2009-1\ has a spectrum with narrow H$\alpha$ emission, [\ion{Ca}{2}] emission, and P-Cygni absorption from many intermediate-mass and Fe-group elements. Perhaps the most distinguishing feature compared to other massive star outbursts is the [\ion{Ca}{2}] emission. In Figure~\ref{f:ugc_spec}, we compare the 15~day spectrum of UGC~2773 OT2009-1\ to spectra of the low-luminosity transients NGC~300 OT2008-1\ \citep{Berger09:ngc} and SN~2008S \citep{Smith09:08s}, as well as SN~IIn~1994W \citep{Chugai04}; all of these objects have [\ion{Ca}{2}] emission in their spectra. \begin{figure} \begin{center} \epsscale{0.9} \rotatebox{90}{ \plotone{u2773_comp.ps}} \caption{Optical spectra of UGC~2773 OT2009-1, NGC~300 OT2008-1\ \citep{Berger09:ngc}, SN~1994W \citep{Chugai04}, and SN~2008S \citep{Smith09:08s}. All spectra have narrow H$\alpha$ and [\ion{Ca}{2}] emission; however, NGC~300 OT2008-1\ and SN~2008S lack the forest of lines (especially \ion{Fe}{2}) that UGC~2773 OT2009-1\ and SN~1994W display.}\label{f:ugc_spec} \end{center} \end{figure} All spectra in Figure~\ref{f:ugc_spec} are relatively similar. The continuum of each spectrum is well-described by a blackbody spectrum, with all four objects having a similar temperature. Each object has a prominent H$\alpha$ emission line, with UGC~2773 OT2009-1\ having a narrower line than the other objects. Additionally, SN~1994W has a strong H$\alpha$ absorption line blueward of its emission peak. NGC~300 OT2008-1\ and SN~2008S are very similar objects with massive (10 -- 25~$M_{\sun}$), dusty progenitors \citep{Prieto08, Berger09:ngc, Bond09}. Their spectra share many characteristics with the yellow hypergiant IRC+10240 \citep{Smith09:08s}. Although UGC~2773 OT2009-1\ shares some spectroscopic properties with these two transients and IRC+10240 (see \citetalias{Smith09:outburst} for additional discussion), the latter objects lack the forest of absorption lines in UGC~2773 OT2009-1. These lines are reminiscent of an F-type supergiant. The P-Cygni profiles of these lines and the hydrogen Balmer emission are very similar to S~Dor during a cool phase \citep[e.g.,][]{Massey00}. SN~1994W was very luminous at peak ($M_{V} \approx -19$~mag), but generated at most 0.03~$M_{\sun}$ of $^{56}$Ni \citep*{Sollerman98}. \citet{Dessart09} presented an alternative method of producing the photometric and spectroscopic properties of this object: the collision of two massive hydrogen shells ejected from the star with no core collapse. Spectra of SNe~IIn are rather heterogeneous (see Figure~5 of \citealt{Smith09:06gy} for a comparison of various objects), and SN~1994W is relatively distinct for its narrow absorption features. Given the spectral similarity between UGC~2773 OT2009-1\ and SN~1994W, the strict upper limit of $^{56}$Ni mass in SN~1994W, and the alternative model of \citet{Dessart09}, one must further question if SN~1994W destroyed its progenitor star. \subsubsection{Contrasting SN~2009ip\ and UGC~2773 OT2009-1} At $t = 0$~days, the temperature of UGC~2773 OT2009-1\ is $\sim\!\!$~ 7000~K (see Section~\ref{ss:sed}), which is similar to the temperature during the ``eruptive'' state of LBVs \citep[e.g.,][]{Humphreys94}. This contrasts with the higher temperature of 10,000~K derived for SN~2009ip\ (see Section~\ref{ss:sed}), which lacks the narrow Fe-group absorption features. Many other LBV giant eruptions have temperatures similar to that of SN~2009ip\ \citep{Humphreys94}. SN~2009ip\ was $\sim\!\!$~ 2~mag brighter at peak than UGC~2773 OT2009-1, and SN~2009ip\ had a much larger increase in luminosity during the year before maximum than UGC~2773 OT2009-1, increasing by $\gtrsim 5$~mag and $\sim\!\!$~ 1~mag over one year, respectively \citepalias{Smith09:outburst}. The fast-moving ejecta of SN~2009ip\ also contrasts with the relatively slow outflow of UGC~2773 OT2009-1. The photometric and spectroscopic differences of these objects suggests different physical mechanisms. Clearly a supersonic explosion is necessary to produce the high-velocity absorption features of SN~2009ip, while UGC~2773 OT2009-1\ shows no indication of an explosion. The differences in temperature and luminosity increase are also indicative of more energy injection (per unit mass) for SN~2009ip. A plausible explanation is that SN~2009ip\ is a LBV giant eruption triggered by an explosion, while UGC~2773 OT2009-1\ is a particularly luminous S~Dor eruption. \subsection{Line Profiles} In this section, we examine the line profiles of H$\alpha$ and Ca lines. These features provide an indication of the kinematics of the emitting material. The narrow lines are a tracer of the pre-shock circumstellar material, while the high-velocity absorption features in the spectra of SN~2009ip\ probe the outburst ejecta. \subsubsection{H$\alpha$}\label{sss:halpha} In Figure~\ref{f:halpha}, we present the H$\alpha$ line profiles of UGC~2773 OT2009-1\ and SN~2009ip. Three separate epochs are shown for each object. Both objects have asymmetric line profiles. There are absorption components at about $-350$km~s$^{-1}$ \ for UGC~2773 OT2009-1\ and between $-3000$ and $-6000$~km~s$^{-1}$ \ for SN~2009ip. The line profile of SN~2009ip\ has a different shape and is much broader than that of UGC~2773 OT2009-1. We have attempted to fit these line profiles, but because of the asymmetry of the profiles, we first fit only the redshifted portion of each line profile and then add an absorption component to reproduce the blueshifted profile. \begin{figure} \begin{center} \epsscale{1.} \rotatebox{90}{ \plotone{halpha.ps}} \caption{Normalized spectra of UGC~2773 OT2009-1\ and SN~2009ip\ near H$\alpha$. The line profiles are fit with Gaussian and Lorentzian profiles, respectively. A profile fit to the redshifted portion of each profile is shown as a dashed line. The grey lines correspond to the redshifted profile with a Gaussian absorption component added. Narrow [\ion{N}{2}] can be seen in the spectrum of UGC~2773 OT2009-1.}\label{f:halpha} \end{center} \end{figure} For the 15~day spectrum of UGC~2773 OT2009-1, we fit a Gaussian with ${\rm FWHM} = 780$~km~s$^{-1}$ \ to the red side of the feature. This value is twice that of the value found by \citetalias{Smith09:outburst} for a spectrum from day 22, but examination of their figures suggest that they reported half-width at half maximum (HWHM) or the standard deviation of the Gaussian fit (which is smaller by a factor of 2.35) rather than FWHM. The 34~day spectrum of UGC~2773 OT2009-1\ is contaminated by host-galaxy emission lines, making a fit to the inner regions of the line profile problematic. Ignoring this region, we were able to fit the redshifted portion of the line profile with a single Gaussian with ${\rm FWHM} = 590$~km~s$^{-1}$ . The 96~day spectrum has lower resolution, but is successfully fit by a Gaussian profile with ${\rm FWHM} = 470$~km~s$^{-1}$ . To account for the asymmetric profile, we add an absorption component to the Gaussian line profiles. Fitting the full profile with two Gaussian functions, the emission component fit to the red side of the line and the absorption component added to fit the blue side of the line, we find absorption minima at $-180$, $-110$, and $-80$~km~s$^{-1}$ \ for the 15, 34, and 96~day spectra, respectively. This is different from the value of the actual minimum ($-350$~km~s$^{-1}$ ) since the relatively strong emission masks the true minimum. The line profiles of the first two spectra (days 4 and 24) of SN~2009ip\ are well fit by Lorentzian profiles with ${\rm FWHM} = 780$~km~s$^{-1}$ \ and the third is best fit by a Lorentzian profile with ${\rm FWHM} = 890$~km~s$^{-1}$ , which are larger than that found by \citetalias{Smith09:outburst}, 550~km~s$^{-1}$ . A Lorentzian profile of 550~km~s$^{-1}$ \ is not a particularly bad fit to our data, but we find that the larger velocities better represent the data. One can also see in Figure~8 of \citetalias{Smith09:outburst}, that the 550~km~s$^{-1}$ \ Lorentzian slightly underpredicts the true FWHM of the line, so the data appear to be consistent. In the 24~day spectrum of SN~2009ip, we see an absorption feature with a minimum at a velocity of about $-3000$~km~s$^{-1}$ . (This high-velocity absorption is seen for all Balmer lines with varying instrument configurations and on two epochs; see Section~\ref{sss:ip_spec}.) This feature is well fit by including a Gaussian absorption component with a minimum at $-2800$~km~s$^{-1}$ . Adding a component with this velocity also improves the fit to the 4~day H$\alpha$ profile slightly, but not in a significant way. The 86~day spectrum shows an even {\it stronger} high-velocity absorption component with the minimum of the absorption at a {\it larger} velocity of $-4800$~km~s$^{-1}$ . The blue wing of the absorption component, representing the fastest moving material, corresponds to a velocity of about $-4500$ and $-7000$~km~s$^{-1}$ \ for the 24 and 86~day spectra, respectively. These velocities are significantly larger than the outflow velocity of 550~km~s$^{-1}$ \ assumed by \citetalias{Smith09:outburst}. They are much larger than the wind speed of LBVs and are larger than the measured velocity for any LBV eruption with the exception of the 1843 eruption of $\eta$~Car, which had some material expelled at 3000 -- 6000~km~s$^{-1}$ \ \citep{Smith08:car}. The velocities measured for SN~2009ip\ are similar to that of the ejecta of typical core-collapse SNe (such as SN~1998S; see Figure~\ref{f:ip_spec} and Section~\ref{sss:ip_spec}) and are somewhat similar to that of Wolf-Rayet winds \citep[e.g.,][]{Abbott87}. We discuss the implications of these features in Section~\ref{ss:exp}. \subsubsection{Permitted and Forbidden \ion{Ca}{2}} Only our first spectrum of SN~2009ip\ covers the \ion{Ca}{2} NIR triplet, and no spectrum shows obvious [\ion{Ca}{2}] $\lambda\lambda 7291$, 7325 lines, similar to the spectra presented by \citetalias{Smith09:outburst}. Furthermore, the Ca H\&K lines are confused by the strong Balmer sequence in SN~2009ip. Because of these factors, it is difficult to evaluate the characteristics of the \ion{Ca}{2} behavior in this object (other than the absent [\ion{Ca}{2}] lines). UGC~2773 OT2009-1, on the other hand, has strong \ion{Ca}{2} features. This can be seen in Figure~\ref{f:opt_spec}. We examine the Ca H\&K, [\ion{Ca}{2}] $\lambda\lambda 7291$, 7325, and \ion{Ca}{2} NIR triplet line profiles in Figure~\ref{f:ca2}. The Ca H\&K lines show a broad absorption extending from $-1000$ to $+500$~km~s$^{-1}$ \ and a minimum at about $-50$~km~s$^{-1}$ \ that does not appear to change significantly between the two epochs. Each component of the \ion{Ca}{2} NIR triplet shows a strong P-Cygni profile with a minimum at approximately $-250$~km~s$^{-1}$ , slightly larger than the minima of Ca H\&K. \begin{figure} \begin{center} \epsscale{3.} \rotatebox{90}{ \plotone{ca2.ps}} \caption{Normalized spectra of UGC~2773 OT2009-1\ near Ca H\&K (top), [\ion{Ca}{2}] $\lambda\lambda 7291$, 7325 (middle), and \ion{Ca}{2} NIR triplet (bottom). Dashed lines indicate the continuum flux and zero velocity for each line. Each member of the multiplet is overplotted for a given spectrum.}\label{f:ca2} \end{center} \end{figure} The [\ion{Ca}{2}] $\lambda\lambda 7291$, 7325 lines are visible in all epochs of our spectroscopy. We confirm the additional line between this doublet seen by \citetalias{Smith09:outburst} and identify this as \ion{Fe}{2} $\lambda 7308$. Both [\ion{Ca}{2}] lines have asymmetric profiles in all spectra; the peak is at zero velocity, but the emission extends further to the red than to the blue. The lines from all epochs have FWHMs of $\sim\!\!$~ 400~km~s$^{-1}$ , which is about half the width of H$\alpha$ (see Section~\ref{sss:halpha}), similar to that found for NGC~300 OT2008-1\ \citep{Berger09:ngc}. \subsection{Spectral Energy Distribution and Dust Emission}\label{ss:sed} Using our available photometry and spectroscopy, we can examine the spectral energy distribution (SED) of both objects. We have only optical spectra of SN~2009ip, which limits our ability to examine multiple blackbody components for this object. A 10,000~K blackbody fits our optical spectra well, which is consistent with that found by \citetalias{Smith09:outburst}. Our single epoch of {\it Swift} photometry occurred during the dramatic fading of the light curve immediately following maximum brightness \citepalias{Smith09:outburst}. In Figure~\ref{f:ipbb}, the {\it Swift} photometry is combined with the unfiltered photometry (approximately $R$ band) of \citetalias{Smith09:outburst} (with an uncertainty of 0.5~mag to account for the 16~hour difference in the epoch of the observations) during the minimum. We overplot the 23~day spectrum for comparison. The optical photometry is consistent with the optical spectrum and a 10,000~K blackbody. The UVW2 flux is also consistent with this blackbody, however, the UVM2 and UVW1 measurements fall well below this curve. Although this may be the result of line blanketing, these data are also consistent with a blackbody curve with a temperature as low as 8000~K. If we ignore the UVW2 measurement, the data can be fit by a 7000~K blackbody. Although our data suggest a possible change in the SED during the fading event, the lack of necessary comparison UV data from a different epoch prevent a clear indication of a change. \begin{figure} \begin{center} \epsscale{0.9} \rotatebox{90}{ \plotone{lbv_bb2.ps}} \caption{UV/Optical photometry of SN~2009ip\ during the fading event immediately after maximum brightness. The blue dotted, orange dashed, and red solid curves correspond to 10,000, 8000, and 7000~K blackbody spectra, respectively. The 23~day spectrum is also plotted to show the consistency with both the photometry and the 10,000~K blackbody. All photometry is consistent with the 8000 and 10,000~K blackbody spectra. Ignoring the bluest (UVW2) filter, the data are also consistent with the 7000~K blackbody.}\label{f:ipbb} \end{center} \end{figure} Using the 15~day optical spectrum and 22~day NIR spectrum, we are able to examine the SED of UGC~2773 OT2009-1\ over nearly a decade in wavelength. Between these dates, the light curve of UGC~2773 OT2009-1\ was essentially constant, having the same magnitude (within 1$\sigma$) \citepalias{Smith09:outburst}. Using the long wavelengths of the NIR spectrum, our data are sensitive to any low-temperature thermal components. We fit a single blackbody to these data, ignoring regions with strong line features and simultaneously fitting the scaling between the optical and NIR spectra. Doing this results in a best-fit temperature of 6800~K. This single blackbody consistently under-predicts the flux at NIR wavelengths. As a result, we have also attempted to fix the spectrum with a double blackbody model. This model, which produces a much better fit, results with temperatures $T_{1} = 6950$~K and $T_{2} = 2100$~K. The full spectrum and associated fits are shown in Figure~\ref{f:ubb}. \begin{figure} \begin{center} \epsscale{0.9} \rotatebox{90}{ \plotone{lbv_bb.ps}} \caption{Optical/NIR spectrum of UGC~2773 OT2009-1. Single (6800~K; dashed orange line) and double (2100 and 6900~K; solid red line) blackbody fits to the spectrum are overplotted. The individual components of the double blackbody fit are shown as blue dotted lines. The double blackbody is a better fit to the data than the single blackbody. Green points show our photometry, which also shows a NIR excess. The $g$-band flux is below that of either curve, but that is likely the result of line blanketing.}\label{f:ubb} \end{center} \end{figure} \citetalias{Smith09:outburst} noted that UGC~2773 OT2009-1\ had a (photometric) NIR excess, but could not distinguish between circumstellar extinction and dust emission. To test the former case, we attempted to fit the spectrum with a single blackbody, but with an additional extinction term. With $R_{V}$ fixed to 3.1, this model did not fit the data well. The model was able to sufficiently reproduce the data if we allowed $R_{V} < 1$, which is unphysical. We therefore conclude that the NIR excess is likely due to dust emission. Scaling our spectrum to our broad-band photometry, we can calibrate the blackbody flux, which in turn constrains the ratio $R/D$, where $R$ is the radius of the blackbody radiation and $D$ is the distance to the object. Using $D = 6 \pm 0.5$~Mpc, we find that the hot and cool blackbodies have radii of $(1.50 \pm 0.16) \times 10^{14}$~cm and $(4.3 \pm 0.4) \times 10^{14}$~cm ($13.0 \pm 1.1$~AU and $29 \pm 2$~AU), respectively. The size of the cool emitting region is of the same order of magnitude of the size of the Homunculus nebula surrounding $\eta$~Car. Following the prescription outlined by \citet*{Smith08:06jc} (and references therein), we can measure the mass of the emitting dust. Specifically, \begin{equation} M_{d} = \frac{400 \pi \rho R_{d}^{2}}{3 T_{d}^{2}}, \end{equation} where $M_{d}$ is the dust mass, $R_{d}$ is the radius of the dust, $T_{d}$ is the dust temperature, and $\rho$ is the dust density. For the values obtained from the spectra, $T_{d} = 2100$~K and $R_{d} = 4.3 \times 10^{14}$~cm, and assuming a dust grain density of $\rho = 2.25$~g~cm$^{-3}$, we find a dust mass of $M_{d} \approx 2 \times 10^{-8} M_{\sun}$. Since there could be a significant amount of dust emission at lower temperatures, this is a lower limit on the total dust mass; however, it is worth noting that this measurement is orders of magnitude less than the dust created in some SNe \citep[e.g.,][and references therein]{Kotak09}. We note depending on the dust composition, the dust temperature may differ from the blackbody temperature by hundreds of degrees. The dust is very close to the star and its temperature is near the limit of grain survival. Given these conditions, it is very likely that pre-existing circumstellar dust was heated and is emitting as it is being vaporized, rather than newly formed dust emitting as it cools. \section{Discussion}\label{s:disc} \subsection{Different Massive Star Outbursts} UGC~2773 OT2009-1\ and SN~2009ip\ provide excellent examples of the diversity of massive star outbursts. UGC~2773 OT2009-1\ increased its optical brightness by $\sim\!\!$~ 1~mag during outburst and has a cool spectrum with many narrow absorption lines and [\ion{Ca}{2}] emission. It occurred near a star cluster containing stars with initial masses of $\sim\!\!$~ $25 M_{\sun}$ and shows evidence for a very cool ($T \approx 2100$~K) thermal component that is radiated by circumstellar dust. SN~2009ip\ had a more massive and relatively isolated progenitor. At peak, it had risen at least 4~mag over the previous year, and its spectrum was hot and dominated by H Balmer emission lines. After having a significant fading and rebrightening over three weeks, it developed high-velocity absorption lines. Its SED is consistent with no dust emission. The high temperature and large increase in optical luminosity for SN~2009ip\ indicates that it was a true giant eruption akin to the 1843 eruption of $\eta$~Car. UGC~2773 OT2009-1, on the other hand, has spectral characteristics similar to that of S~Dor at maximum. The relatively small increase in optical luminosity may indicate that UGC~2773 OT2009-1\ was the result of normal S~Dor variability, but that it was a particularly luminous maximum. While the largest normal variation of S~Dor stars vary by $\sim\!\!$~ 3~mag in the optical \citep{vanGenderen01}, UGC~2773 OT2009-1\ has varied by at least $\sim\!\!$~ 5~mag (including the current eruption) over the last ten years (Section~\ref{s:obs}; see also \citetalias{Smith09:outburst}). The outbursts of NGC~300 OT2008-1\ and SN~2008S both had relatively low temperature SEDs \citep{Berger09:ngc}, but neither had the forest of absorption lines found in UGC~2773 OT2009-1. All three objects had [\ion{Ca}{2}] emission, but as suggested by \citetalias{Smith09:outburst}, this may be linked to the circumstellar environment, and particularly dust destruction, rather than the event. Our observations have shown that there is dust in the circumstellar environment of the progenitor, and that it was likely pre-existing dust that is in the process of being vaporized. Additionally, SN~1999bw had [\ion{Ca}{2}] emission \citep{Garnavich99} and had an IR excess consistent with dust emission at late times \citep{Sugerman04}. All four massive star outbursts with observed [\ion{Ca}{2}] emission (SN~1999bw, SN~2008S, NGC~300 OT2008-1, and UGC~2773 OT2009-1) have evidence of circumstellar dust. We do note that SN~2000ch had a $\sim\!\!$~ 7000~K spectrum and an infrared excess consistent with dust emission, but no strong [\ion{Ca}{2}] emission was detected in the relatively low signal-to-noise spectra presented by \citet{Wagner04}. It is therefore possible to have a cool object and circumstellar dust yet not have [\ion{Ca}{2}] emission. Although UGC~2773 OT2009-1, NGC~300 OT2008-1, and SN~2008S have circumstellar dust and similar temperatures, other than the narrow H Balmer and [\ion{Ca}{2}] (which is linked to the presence of circumstellar dust) emission, the spectra and progenitors are not particularly similar. Particularly, NGC~300 OT2008-1\ and SN~2008S had relatively featureless spectra and progenitors with initial masses of 10 -- $25 M_{\sun}$, while UGC~2773 OT2009-1\ had spectrum dominated by narrow lines and a more massive progenitor ($\gtrsim 25 M_{\sun}$). Additional data are necessary to determine if the outburst mechanisms in these objects are similar. Using SN~2009ip\ and UGC~2773 OT2009-1\ as examples, there appears to be two distinct elements that determine the observational properties of massive star outbursts. The first is the temperature of the outburst, which may be related to the increase in luminosity, the instability that causes the eruption, the width of the emission lines, and possibly the energetics of the outburst and if there is an explosion (see Section~\ref{ss:exp}). This directly determines the shape of the optical SED, the ionization, if there is a forest of absorption lines, and possibly the shape of the line profiles, if there is high-velocity absorption. The other characteristic is the amount of circumstellar dust, which may cause strong Ca emission (and particularly [\ion{Ca}{2}] emission) and will determine the shape of the SED at longer wavelengths. UGC~2773 OT2009-1\ and SN~2009ip\ would occupy very different regions of the parameter space created by these two dimensions. NGC~300 OT2008-1\ and SN~2008S would be close to UGC~2773 OT2009-1, while LBV giant eruptions such as SN~1997bs would be close to SN~2009ip. It remains to be seen if there are hot massive star outbursts with a large amount of circumstellar dust or if there are cool massive star outbursts with little circumstellar dust. $\eta$~Car has $0.125 M_{\sun}$ of dust surrounding it \citep{Smith03}; if it were to have another giant eruption today, would it be cool? SN~1999bw had [\ion{Ca}{2}] emission and displayed dust emission at late times; was it hot? An IR survey of recent massive star outbursts with good spectroscopic coverage may provide these answers. In the future, optical and NIR observations may be sufficient to determine these characteristics for other massive star outbursts. \subsection{SN~2009ip: A Supersonic Explosion}\label{ss:exp} The spectra of SN~2009ip\ have absorption attributed to high-velocity (up to $\sim\!\!$~ 7000~km~s$^{-1}$ ) material (see Section~\ref{sss:halpha}). Contrary to what is expected from a single outburst or explosion, the velocity of the absorption feature increases with time. In a typical SN, the ejecta naturally follow a Hubble law with the highest velocity material being the most distant from the explosion site. Spectral lines have a blueshifted absorption due to the scattering processes in the photosphere of the SN. Low velocity material is hidden behind the photosphere, only to be revealed at later times. As the photosphere recedes, the highest-velocity material becomes optically thin, resulting in the blueshifted velocity of a spectral line to decrease with time. Since the absorbing material must be at just slightly larger radii than the photosphere, the high-velocity material must have been ejected during the eruption. (If the absorbing material were from a previous eruption, the ejecta from the more recent eruption would have had to be moving even faster.) It is possible that the high-velocity absorption is a component of P-Cygni features from the ejected material. The Lorentzian profile slightly underestimates the emission flux in the 86~day spectrum (see Figure~\ref{f:halpha}), which may be the result of P-Cygni emission contributing to the line. Since the high-velocity absorption is coming from the ejecta, the outburst of SN~2009ip\ must have been extremely energetic, expelling a large amount of material at very high velocities. However, for the velocity to {\it increase} with time, either the ejecta must not follow a Hubble expansion or the radius of the photosphere (in velocity space) must somehow increase with time. In a single explosion, the ejecta naturally follow a Hubble law; however, multiple explosions can change the velocity profile of the ejecta. If two explosions occurred in short succession, one can produce the inverted velocity gradient seen in SN~2009ip. In this toy model, the photosphere would recede into the ejecta of the first explosion, but at some point the fastest-moving ejecta of the second explosion would overtake the photosphere, increasing the velocity. If there are no other explosions, the velocity of the absorption would decrease from there. The photometric behavior of SN~2009ip\ is consistent with this picture. The first explosion would produce the fast rise to maximum. As noted by \citetalias{Smith09:outburst}, the timescale of the fading is much shorter than the timescales for many physical processes such as dust extinction.\footnote{The calculation by \citepalias{Smith09:outburst} for the time until dust formation for SN~2009ip\ assumes a velocity of 500~km~s$^{-1}$ . Although the ejecta are moving much faster than this assumed value, they would need to have a velocity of $\gtrsim 20,000$~km~s$^{-1}$ \ to reach the sublimation radius at the time of fading.} This behavior is very similar to that of SN~2000ch, which brightened by 2.1~mag in 9~days to maximum, then immediately faded by 3.4~mag in 7~days, immediately followed by a 2.2~mag rise in 4~days, after which the magnitude stayed relatively constant \citep{Wagner04}. Spectra of SN~2000ch taken during the fading and on its plateau show no strong evidence for high-velocity ejecta, but the spectra may not be of high enough quality to see these features. \citetalias{Smith09:outburst} hypothesized that rapid fading may have been caused by an optically thick shell being ejected after the first outburst. If this process did occur in SN~2009ip, then there are several implications: (1) the velocity of the absorption should eventually decrease, (2) the interaction of the ejecta from the two explosions could be a significant source of X-ray and radio emission, and (3) the X-rays might excite certain elements producing high-excitation lines such as \ion{He}{2} in optical spectra. Our X-ray limit of $L_{X} < 4.8 \times 10^{38}$~ergs~s$^{-1}$ taken during the minimum is not particularly constraining. We do not detect any \ion{He}{2} $\lambda 4686$ emission in the 4 or 24~day spectra; however, there is a low significance detection of a line consistent with \ion{He}{2} $\lambda 4686$ emission in the 86~day spectrum. Additional spectroscopy is necessary to determine the late-time velocity gradient. \section{Conclusions}\label{s:con} We have presented extensive UV, optical, and NIR data for two transients, SN~2009ip\ and UGC~2773 OT2009-1. Although these events appear to be similar phenomena (luminous outbursts of massive stars), the details of the events show that there are many differences. These differences provide examples of the diversity of such events. A previous study of these events, \citetalias{Smith09:outburst}, provided an initial analysis of the object. Although the two studies agree on many points, our interpretation of the entire data set is somewhat different than that of \citetalias{Smith09:outburst}. In particular, we agree that based on pre-outburst {\it HST} imaging, historical light curves, and outburst spectroscopy, the progenitors of SN~2009ip\ and UGC~2773 OT2009-1\ are LBVs with masses of $\gtrsim 60$ and $\gtrsim 25 M_{\sun}$, respectively. We also agree that the spectra of the two events are significantly different, but consistent with known LBVs or LBV outbursts. While UGC~2773 OT2009-1\ had a cooler spectrum with a forest of absorption lines reminiscent of a F-type supergiant (similar to S~Dor in its high state), SN~2009ip\ had a hot spectrum and exhibited mainly H Balmer emission (similar to other LBV giant eruptions). The spectral characteristics (particularly [\ion{Ca}{2}] emission) and circumstellar dust link UGC~2773 OT2009-1\ to the lower-mass, dust-obscured progenitors of NGC~300 OT2008-1\ and SN~2008S. We agree that the progenitors of these objects are all massive stars and may have many characteristics similar to those of the LBV class, which could extend the mass range for LBV-like activity to relatively low-mass stars. However, there are distinct differences between the analyses of \citetalias{Smith09:outburst} and of that presented here. Specifically, the initial mass ranges for the progenitors are slightly different in the two studies, with \citetalias{Smith09:outburst} estimating 50 -- $80 M_{\sun}$ (instead of $\gtrsim 60 M_{\sun}$) and $\gtrsim 20 M_{\sun}$ (instead of $\gtrsim 25 M_{\sun}$) for SN~2009ip\ and UGC~2773 OT2009-1, respectively. The differences lie in the conversion from \protect\hbox{\it HST} \ filters to Bessell filters and the adapted color range for the progenitor of SN~2009ip, and the additional information provided by stars in the vicinity of the progenitor of UGC~2773 OT2009-1. Our interpretation of the exact nature of UGC~2773 OT2009-1\ differs from that of \citetalias{Smith09:outburst}. While \citetalias{Smith09:outburst} contends that this object is a true giant eruption of an LBV, we question this assertion and propose that it may be the result of extreme S~Dor variability. While \citetalias{Smith09:outburst} found an NIR excess for UGC~2773 OT2009-1, hypothesizing that there may be dust emission as it is vaporized, we find more conclusive evidence for this scenario through our NIR spectroscopy. The NIR spectrum is consistent with an additional blackbody with $T \approx 2100$~K, but is inconsistent with reddening by dust. The presence of this dust indicates that the initial mass estimate of the progenitor (based on optical {\it HST} imaging) is a lower limit, and that the true initial mass is likely much larger. In addition to what was discussed by \citetalias{Smith09:outburst}, we have also detected high-velocity absorption in the spectra of SN~2009ip, indicative of an explosion (as opposed to subsonic outburst). The absorption has an inverse velocity gradient suggesting multiple explosions in short succession. The rapid fading and brightening shortly after maximum brightness noted by \citepalias{Smith09:outburst} is consistent with multiple explosions, where a second explosion ejects an optically thick shell that temporarily dims the object. We also note the spectroscopic similarity of UGC~2773 OT2009-1\ and SN~1994W, which \citet{Dessart09} has previously suggested was not a true SN that destroyed its progenitor star. SN~2009ip\ and UGC~2773 OT2009-1\ are very different manifestations of a similar phenomenon: extreme brightening of massive stars. With these objects and similar events (such as $\eta$~Car, NGC~300 OT2008-1, and SNe~1961V, 1954J, 1997bs, 2000ch, 2002kg, and 2008S), we show that luminous outbursts of massive stars are very heterogeneous. Some of this diversity is likely linked to the instability that causes the eruption, while some is caused by the circumstellar environment. Additional observations of new massive star eruptions are necessary to determine the physical mechanisms of the eruptions, the content of the circumstellar environments, and whether the two are physically connected. \begin{acknowledgments} {\it Facilities:} \facility{ARC (TripleSpec), FMO:31in (FanCam), Gemini:Gillett (GMOS), HST (WFPC2), Magellan:Baade (IMACS), Magellan:Clay (MagE), MMT (Blue Channel), Swift (UVOT, XRT)} \bigskip R.J.F.\ is supported by a Clay Fellowship. O.D.F.\ is grateful for support from NASA GSRP, ARCS, and VSGC. E.M.L.\ is supported in part by a Ford Foundation Predoctoral Fellowship. We are indebted to the staffs at the APO, Gemini, Magellan, and MMT Observatories for their dedicated services. We thank K.\ Olsen and R.\ McDermid for obtaining some of the data presented in the paper. We thank R.\ Chornock and R.\ Kirshner for stimulating discussions about the transients. Based on observations made with the NASA/ESA Hubble Space Telescope, obtained from the data archive at the Space Telescope Science Institute. STScI is operated by the Association of Universities for Research in Astronomy, Inc.\ under NASA contract NAS 5-26555. Based in part on observations obtained at the Gemini Observatory, which is operated by the Association of Universities for Research in Astronomy, Inc., under a cooperative agreement with the US National Science Foundation on behalf of the Gemini partnership: the NSF (United States), the Science and Technology Facilities Council (United Kingdom), the National Research Council (Canada), CONICYT (Chile), the Australian Research Council (Australia), Minist\'{e}rio da Ci\^{e}ncia e Tecnologia (Brazil) and Ministerio de Ciencia, Tecnolog\'{i}a e Innovaci\'{o}n Productiva (Argentina); the 6.5 meter Magellan Telescopes located at Las Campanas Observatory, Chile; the MMT Observatory, a joint facility of the Smithsonian Institution and the University of Arizona; the Fan Mountain Observatory 0.8 meter telescope; and the Apache Point Observatory 3.5 meter telescope, which is owned and operated by the Astrophysical Research Consortium. We acknowledge the use of public data from the {\it Swift} data archive. \end{acknowledgments} \bibliographystyle{fapj}
\section{Introduction} Writing safe and robust code is a hard task; writing safe and robust multi-threaded low-level code is even harder. In this paper we present a minimal, low-level concurrent language with advanced region-based memory management and hierarchical lock-based synchronization primitives. Region-based memory management achieves efficiency by bulk allocation and deallocation of objects in segments of memory called regions. Similar to other approaches, our regions are organized in a hierarchical manner such that each region is physically allocated within a single parent region and may contain multiple child regions. This hierarchical structure imposes an ownership relation as well as lifetime constraints over regions. Unlike other languages employing hierarchical regions, our language allows early subtree deallocation in the presence of region sharing between threads. In addition, each thread is obliged to release each region it owns by the end of its scope. Multi-threaded programs that interact through shared memory generate random execution interleavings. A data race occurs in a multi-threaded program when there exists an interleaving such that some thread accesses a memory location while some other thread attempts to write to it. So far, type systems and analyses that guarantee race freedom~\cite{FlanaganAbadi@CONCUR-99} have mainly focused on lexically-scoped constructs. The key idea in those systems is to statically track or infer the lockset held at each program point. In the language presented in this paper, implicit reentrant locks are used to protect regions from data races. Our locking primitives are non-lexically scoped. Locks also follow the hierarchical structure of regions so that each region is protected by its own lock as well as the locks of all its ancestors. Furthermore, our language allows regions and locks to be safely aliased, escape the lexical scope when passed to a new thread, or become logically separated from the remaining hierarchy. These features are invaluable for expressing numerous idioms of multi-threaded programming such as \emph{sharing}, \emph{region ownership} or \emph{lock ownership transfers}, \emph{thread-local regions} and \emph{region migration}. \section{Language Design} \label{sec:goals} We briefly outline the main design goals for our language, as well as some of the main design decisions that we made to serve these goals. \myparagraph{Low-level and concurrent} Our language must efficiently support systems programming. As such, it should cater for memory management and concurrency. It also needs to be low-level: it is not intended to be used by programmers but as a target language of higher-level systems programming languages. \myparagraph{Static safety guarantees} We define safety in terms of \emph{memory safety} and absence of \emph{data races}. A static type system should guarantee that well-typed programs are safe, with minimal run-time overhead. \myparagraph{Safe region-based memory management} Similarly to other languages for safe systems programming (e.g.\ Cyclone) our language employs region-based memory management, which achieves efficiency by \emph{bulk allocation} and \emph{deallocation} of objects in segments of memory (\emph{regions}). Statically typed regions~\cite{TofteTalpin@POPL-94,Capabilities@TOPLAS-00} guarantee the absence of dangling pointer dereferences, multiple release operations of the same memory area, and memory leaks. Traditional stack-based regions~\cite{TofteTalpin@POPL-94} are limiting as they cannot be deallocated early. Furthermore, the stack-based discipline fails to model region lifetimes in concurrent languages, where the lifetime of a shared region depends on the lifetime of the longest-lived thread accessing that region. In contrast, we want regions that can be \emph{deallocated early} and that can safely be \emph{shared} between concurrent threads. We opt for a \emph{hierarchical region}~\cite{GayAiken@PLDI-01} organization: each region is physically allocated within a single parent region and may contain multiple child regions. Early region deallocation in our multi-level hierarchy automatically deallocates the immediate subtree of a region without having to deallocate each region of the subtree recursively. The hierarchical region structure imposes the constraint that a child region is \emph{live} only when its ancestors are live. In order to allow a function to access a region without having to pass all its ancestors explicitly, we allow ancestors to be abstracted (i.e., our language supports \emph{hierarchy abstraction}) for the duration of the function call. To maintain the \emph{liveness} invariant we require that the abstracted parents are \emph{live} before and after the call. Regions whose parent information has been abstracted cannot be passed to a new thread as this may be unsound. \myparagraph{Race freedom} To prevent data races we use \emph{lock-based} mutual exclusion. Instead of having a separate mechanism for locks, we opt for a uniform treatment of locks and regions: locks are placed in the same hierarchy as regions and enjoy similar properties. Each region is protected by its own private lock and by the locks of its ancestors. The semantics of region locking is that the entire subtree of a region is \emph{atomically locked} once the lock for that region has been acquired. Hierarchical locking can model complex synchronization strategies and lifts the burden of having to deal with explicit acquisition of multiple locks. Although deadlocks are possible, they can be \emph{avoided} by acquiring a single lock for a group of regions rather than acquiring multiple locks for each region separately. Additionally, our language provides explicit locking primitives, which in turn allow a higher degree of concurrency than lexically-scoped locking, as some locks can be released early. \myparagraph{Region polymorphism and aliasing} Our language supports \emph{region polymorphism}: it is possible to pass regions as parameters to functions or concurrent threads. This enables \emph{region aliasing}: one actual region could be passed in the place of two distinct formal region parameters. In the presence of mutual exclusion and early region deallocation, aliasing is dangerous. Our language allows safe region aliasing with minimal restrictions. The mechanism that we employ for this purpose also allows us to encode numerous useful idioms of concurrent programming, such as \emph{region migration}, \emph{lock ownership transfers}, \emph{region sharing}, and \emph{thread-local regions}. \section{Language Features through Examples} \label{sec:examples} Our regions are lexically-scoped first-class citizens; they are manipulated via explicit handles. For instance, a region handle can be used for releasing a region early, for allocating references and regions within it, or for locking it. Our language uses a \emph{type and effect system} to guarantee that regions and their contents are properly used. The details will be made clear in Sections~\ref{sec:language} and~\ref{sec:typing}. Here, we present the main features of our language through examples. We try to avoid technical issues as much as possible; however, some characteristics of the type and effect system are revealed in this section and their presence is justified. Furthermore, to simplify the presentation in this section, we use abbreviations for a few language constructs that we expect the readers will find more intuitive. \begin{example}[Simple Region Usage]\normalfont \newcommand\excom[2][8cm]{\nabstab{-#1}{/\!/}\hspace{0.5em}\textit{#2}} This example shows a typical region use. New regions are allocated via the $\cfont{newrgn}$ construct. This construct requires a handle to an existing region ($\mathit{heap}$ in this case), in which the new region will be allocated, and introduces a type-level name ($\rho$) and a fresh handle ($h$) for the new region. The handle $h$ is then used to allocate a new integer in region $\rho$; a reference to this integer ($z$) is created. Finally, the region is deallocated before the end of its lexical scope. \begin{ndisplay} \cfont{newrgn} \ \nemphthree{\rho}, \nemphtwo{h} \ \cfont{at} \ \nemphtwo{\mathit{heap}} \ \cfont{in} \excom{$\nemphthree{\{\conelt{\rho}{1,1}{\rho_H}\}}$} \\ \ntab[1] \cfont{let} \ z \ \cfont{=} \ \cfont{new} \ 10 \ \cfont{at} \ \nemphtwo{h} \ \cfont{in} \\ \ntab[2] \ldots \\ \ntab[2] z \ \cfont{:=} \ \cfont{deref} \ z \cdef{+} 5 \ \cfont{;} \\ \ntab[2] \ldots \\ \ntab[2] \cfont{free} \ \nemphtwo{h} \ \cfont{;} \excom{$\nemphthree{\{\,\}}$ \hspace{1em} --- empty effect, $\rho$ is no longer alive} \\ \ntab[2] \ldots \end{ndisplay} The comments on the right-hand side of the example's code show the current \emph{effect}. An effect is roughly a set of \emph{capabilities} that are held at a given program point. Right after creation of region $\rho$, the entry $\conelt{\rho}{1,1}{\rho_H}$ is added to the effect; this means that a capability (``$1,1$'' --- we will later explain what this means) is held for region $\rho$, which resides in the heap region ($\rho_H$). Regions start their life as local to a thread and their contents can be directly accessed. For instance, a reference $z$ can be created in $\rho$, dereferenced and assigned a new value, as long as the type system can verify that a proper capability for $\rho$ is present in the current effect. Deallocation of $\rho$ removes the capability from the effect; once that is done, the region's contents become inaccessible. \end{example} \begin{example}[Hierarchical Regions]\normalfont \newcommand\excom[2][6cm]{\nabstab{-#1}{/\!/}\hspace{0.5em}\textit{#2}} In the previous example a trivial hierarchy was created by allocating region $\rho$ within the $heap$ region. It is possible to construct richer region hierarchies. As in the previous example, the code below allocates a new region $\rho_1$ within the heap. Other regions can be then allocated within $\rho_1$, e.g.\ $\rho_2$; this can done by passing the handle of $\rho_1$ to the region creation construct. Similarly, regions $\rho_3$ and $\rho_4$ can be allocated within region $\rho_2$. \begin{flushleft} \begin{minipage}{12.5cm}% \begin{ndisplay} \cfont{newrgn} \ \nemphthree{\rho_1},\nemphtwo{h} \ \cfont{at} \ \nemphtwo{\mathit{heap}} \ \cfont{in} \excom{$\nemphthree{\{\conelt{\rho_1}{1,1}{\rho_H}\}}$} \\ \ntab[1] \ldots \\ \ntab[1] \cfont{newrgn} \ \nemphthree{\rho_2},\nemphtwo{h_2} \ \cfont{at} \ \nemphtwo{h_1} \ \cfont{in} \excom{$\nemphthree{\{\conelt{\rho_1}{1,1}{\rho_H}, \ \conelt{\rho_2}{1,1}{\rho_1}\}}$} \\ \ntab[2] \ldots \\ \ntab[2] \cfont{newrgn} \ \nemphthree{\rho_3},\nemphtwo{h_3} \ \cfont{at} \ \nemphtwo{h_2} \ \cfont{in} \excom{$\nemphthree{\{\conelt{\rho_1}{1,1}{\rho_H}, \ \conelt{\rho_2}{1,1}{\rho_1}, \ \conelt{\rho_3}{1,1}{\rho_2}\}}$} \\ \ntab[3] \cfont{newrgn} \ \nemphthree{\rho_4},\nemphtwo{h_4} \ \cfont{at} \ \nemphtwo{h_2} \ \cfont{in} \excom{$\nemphthree{\{\conelt{\rho_1}{1,1}{\rho_H}, \ \conelt{\rho_2}{1,1}{\rho_1}, \ \conelt{\rho_3}{1,1}{\rho_2}, \ \conelt{\rho_4}{1,1}{\rho_2}\}}$} \\ \ntab[4] \ldots \end{ndisplay} \end{minipage}% \hfill \begin{minipage}{2.5cm}% \begin{tikzpicture}[node distance=1.5cm][->,semithick] \tikzstyle{every state}=[scale=0.75,draw=none,text=black] \node[state] (H) {$\rho_H^{1,0}$}; \node[state] (D) [below of=H] {$\rho_1^{1,1}$}; \path (H) edge (D); \node[state] (A) [below of=D] {$\rho_2^{1,1}$}; \path (D) edge (A); \node[state] (F) [below left of=D] {$\dots$}; \path (D) edge (F); \node[state] (B) [below left of=A] {$\rho_3^{1,1}$}; \path (A) edge (B); \node[state] (C) [below right of=A] {$\rho_4^{1,1}$}; \path (A) edge (C); \node[state] (E) [below right of=D] {$\dots$}; \path (D) edge (E); \end{tikzpicture} \end{minipage} \end{flushleft} Our language allows regions to be allocated at any level of the hierarchy. For instance, it is possible to allocate more regions within region $\rho_1$, in the lexical scope of region $\rho_4$. \end{example} \begin{example}[Bulk Region Deallocation]\normalfont \newcommand\excom[2][6cm]{\nabstab{-#1}{/\!/}\hspace{0.5em}\textit{#2}} In the first example a single region was deallocated. That region was a \emph{leaf} node in the hierarchy; it contained no sub-regions. In the general case, when a region is deallocated, the entire subtree below that region is also deallocated. This is what happens if, in the code of the previous example, we deallocate region $\rho_2$ within the innermost scope; regions $\rho_3$ and $\rho_4$ are also deallocated. They are all removed from the current effect and thus are no longer accessible. \begin{flushleft} \begin{minipage}{12.5cm}% \begin{ndisplay} \cfont{newrgn} \ \nemphthree{\rho_1},\nemphtwo{h} \ \cfont{at} \ \nemphtwo{\mathit{heap}} \ \cfont{in} \excom{$\nemphthree{\{\conelt{\rho_1}{1,1}{\rho_H}\}}$} \\ \ntab[1] \ldots \\ \ntab[1] \cfont{newrgn} \ \nemphthree{\rho_2},\nemphtwo{h_2} \ \cfont{at} \ \nemphtwo{h_1} \ \cfont{in} \excom{$\nemphthree{\{\conelt{\rho_1}{1,1}{\rho_H}, \ \conelt{\rho_2}{1,1}{\rho_1}\}}$} \\ \ntab[2] \ldots \\ \ntab[2] \cfont{newrgn} \ \nemphthree{\rho_3},\nemphtwo{h_3} \ \cfont{at} \ \nemphtwo{h_2} \ \cfont{in} \excom{$\nemphthree{\{\conelt{\rho_1}{1,1}{\rho_H}, \ \conelt{\rho_2}{1,1}{\rho_1}, \ \conelt{\rho_3}{1,1}{\rho_2}\}}$} \\ \ntab[3] \cfont{newrgn} \ \nemphthree{\rho_4},\nemphtwo{h_4} \ \cfont{at} \ \nemphtwo{h_2} \ \cfont{in} \excom{$\nemphthree{\{\conelt{\rho_1}{1,1}{\rho_H}, \ \conelt{\rho_2}{1,1}{\rho_1}, \ \conelt{\rho_3}{1,1}{\rho_2}, \ \conelt{\rho_4}{1,1}{\rho_2}\}}$} \\ \ntab[4] \ldots \\ \ntab[4] \cfont{free} \ \nemphtwo{h_2}; \excom{$\nemphthree{\{\conelt{\rho_1}{1,1}{\rho_H}\}}$} \\ \ntab[4] \ldots \excom{$\rho_2$, $\rho_3$ and $\rho_4$ are no longer alive} \end{ndisplay} \end{minipage}% \hfill \begin{minipage}{2.5cm}% \begin{tikzpicture}[node distance=1.5cm][->,semithick] \tikzstyle{every state}=[scale=0.75,draw=none,text=black] \node[state] (H) {$\rho_H^{1,0}$}; \node[state] (D) [below of=H] {$\rho_1^{1,1}$}; \path (H) edge (D); \node[state] (A) [below of=D] {$\rho_2^{1,1}$}; \path (D) edge (A); \node[state] (F) [below left of=D] {$\dots$}; \path (D) edge (F); \node[state] (B) [below left of=A] {$\rho_3^{1,1}$}; \path (A) edge (B); \node[state] (C) [below right of=A] {$\rho_4^{1,1}$}; \path (A) edge (C); \node[state] (E) [below right of=D] {$\dots$}; \path (D) edge (E); \begin{scope}[color=gray,line width=4pt] \draw (-0.5,-2.2) -- (0.5,-3.2); \draw (-0.5,-3.2) -- (0.5,-2.2); \end{scope} \end{tikzpicture} \end{minipage}% \end{flushleft} \end{example} \begin{example}[Region Migration]\normalfont \newcommand\excom[2][8cm]{\nabstab{-#1}{/\!/}\hspace{0.5em}\textit{#2}} A common multi-threaded programming idiom is to use \emph{thread-local} data. At any time, only one thread will have access to such data and therefore no locking is required. A thread can transfer thread-local data to another thread but, doing so, it loses access to the data. This idiom is known as \emph{migration}. Our language encodes thread-local data and data migration. As we have seen, newly created regions are considered thread-local; a capability for them is added to the current effect. We support data migration by allowing such capabilities to be transferred to other threads. The following example illustrates region migration. A server thread is defined, which executes an infinite loop. In every iteration, a new region is created and is initialized with client data. The contents of the region are then processed and finally transferred to a newly created (spawned) thread. \begin{ndisplay} \cfont{def} \ \cfont{server} = \Lambda \nemphthree{\rho_H}. \ \lambda \nemphtwo{\mathit{heap}}. \\ \ntab[1] \cfont{while} \ (\cfont{true}) \ \cfont{do} \\ \ntab[2] \cfont{newrgn} \ \nemphthree{\rho},\nemphtwo{h} \ \cfont{at} \ \nemphtwo{\mathit{heap}} \ \cfont{in} \excom{$\nemphthree{\{\conelt{\rho}{1,1}{\rho_H}\}}$} \\ \ntab[3] \cfont{let} \ z = \cfont{wait\_data}[\nemphthree{\rho}](\nemphtwo{h}) \ \cfont{in} \excom{region $\rho$ is thread-local} \\ \ntab[4] \cfont{process}(z); \\ \ntab[4] \cfont{spawn} \ \cfont{output}[\nemphthree{\rho}](\nemphtwo{h}, z); \excom{$\nemphthree{\{\,\}}$ \hspace{1em} --- empty effect, $\rho$ migrates to $\cfont{output}$} \\ \ntab[4] \ldots \excom{$\rho$ cannot be accessed here} \end{ndisplay} The server thread accepts the heap region and its handle. Within the infinite loop, it allocates a new region $\rho$ in the heap. Its handle $h$ is passed to function $\cfont{wait\_data}$, which is supposed to fill the region $\rho$ with client data ($z$). Function $\cfont{process}$ is then called and works on the data. Until this point, region $\rho$ is thread-local and accessible to the server thread, so no explicit locking is required. Now, let us assume that we want the processed data to be output by a different thread, e.g.\ to avoid an unnecessary delay on the server thread. A new thread $\cfont{output}$ is spawned and receives the region handle $h$ and the reference $z$ to the client data. The capability $\conelt{\rho}{1,1}{\rho_H}$ is removed from the effect of $\cfont{server}$ and is added to the input effect of thread $\cfont{output}$. Therefore, region $\rho$ has now become thread-local to thread $\cfont{output}$, which can access it directly, while it is no longer accessible to the server thread. \end{example} \begin{example}[Region Sharing]\normalfont \newcommand\excom[2][8cm]{\nabstab{-#1}{/\!/}\hspace{0.5em}\textit{#2}} In the previous examples, capabilities for all regions were ``$1,1$'' which, as we roughly explained, corresponds to thread-local. In general, a capability for a region consists of two natural numbers; the first denotes the \emph{region count}, whereas the second denotes the \emph{lock count}. When the region count is positive, the region is definitely alive. Similarly, when the lock count is positive, memory accesses to this region's contents are guaranteed to be race free. Capabilities with counts other than 1 can be used for \emph{sharing} regions between threads. Multithreaded programs often share data for communication purposes. In this example, a server thread almost identical to that of the previous example is defined. The programmer's intention here, however, is to process the data and display it in parallel. Therefore, the $\cfont{output}$ thread is spawned first and then the server thread starts processing the data. \begin{ndisplay} \cfont{def} \ \cfont{server} = \Lambda \nemphthree{\rho_H}. \ \lambda \nemphtwo{\mathit{heap}}. \\ \ntab[1] \cfont{while} \ (\cfont{true}) \ \cfont{do} \\ \ntab[2] \cfont{newrgn} \ \nemphthree{\rho},\nemphtwo{h} \ \cfont{at} \ \nemphtwo{\mathit{heap}} \ \cfont{in} \excom{$\{\conelt{\rho}{1,1}{\rho_H}\}$} \\ \ntab[3] \cfont{let} \ z = \cfont{wait\_data}[\nemphthree{\rho}](\nemphtwo{h}) \ \cfont{in} \\ \ntab[4] \cfont{share} \ \nemphtwo{h}; \ \cfont{unlock} \ \nemphtwo{h}; \excom{\{\conelt{\rho}{2,0}{\rho_H}\}} \\ \ntab[4] \cfont{spawn} \ \cfont{output}[\nemphthree{\rho}](\nemphtwo{h}, z); \excom{$\{\conelt{\rho}{1,0}{\rho_H}\}$ \hspace{1em} --- $\cfont{output}$ consumes $\conelt{\rho}{1,0}{\rho_H}$} \\ \ntab[4] \cfont{while} \ (\cfont{!}\,\mathit{finished}) \ \cfont{do} \\ \ntab[5] \cfont{lock} \ \nemphtwo{h}; \excom{\{\conelt{\rho}{1,1}{\rho_H}\}} \\ \ntab[5] \cfont{process}(z); \\ \ntab[5] \cfont{unlock} \ \nemphtwo{h} \excom{\{\conelt{\rho}{1,0}{\rho_H}\}} \end{ndisplay} Operator $\cfont{share}$ increases the region count and operator $\cfont{unlock}$ decreases the lock count. As a consequence, starting with capability $\conelt{\rho}{1,1}{\rho_H}$, we end up with $\conelt{\rho}{2,0}{\rho_H}$. When $\cfont{output}$ is spawned, it consumes ``half'' of this capability ($\conelt{\rho}{1,0}{\rho_H}$); the remaining ``half'' ($\conelt{\rho}{1,0}{\rho_H}$) is still held by the server thread. Region $\rho$ is now shared between the two threads; however, none of them can access its data directly, as this may lead to a data race. The $\cfont{lock}$ and $\cfont{unlock}$ operators have to be used for explicitly locking and unlocking the region, before safely accessing its contents. Processing is now performed iteratively; the server thread avoids locking the region for long periods of time, thus allowing the $\cfont{output}$ thread to execute a similar loop and gain access to the region when needed. \end{example} \begin{example}[Hierarchical Locking]\normalfont \newcommand\excom[2][8cm]{\nabstab{-#1}{/\!/}\hspace{0.5em}\textit{#2}} In the previous example, locking and unlocking was performed on a leaf region. In general, locking a region in the hierarchy has the effect of atomically locking its subregions as well. A region is accessible when it has been locked by the current thread or when at least one of its ancestors has been locked. Hierarchical locking can be useful when a set of locks needs to be acquired atomically. In this example, we assume that two hash tables ($\mathit{tbl}_1$ and $\mathit{tbl}_2$) are used. An object with a given key must be removed from $\mathit{tbl}_1$, which resides in region $\rho_1$, and must be inserted in $\mathit{tbl}_2$, which resides in region $\rho_2$. We can atomically acquire access to both regions $\rho_1$ and $\rho_2$, by locking a common ancestor of theirs. \begin{ndisplay}[indent=0.5cm] \cfont{lock}\ \nemphtwo{h}; \excom{the handle of a common ancestor of $\rho_1$ and $\rho_2$} \\ \ntab[1] \cfont{let} \ obj = \cfont{hash\_remove}[\nemphthree{\rho_1}](tbl_1, key) \ \cfont{in} \\ \ntab[2] \cfont{hash\_insert}[\nemphthree{\rho_2}](tbl_2, key, obj); \\ \cfont{unlock}\ \nemphtwo{h} \end{ndisplay} \end{example} \begin{example}[Region Aliasing]\normalfont \newcommand\excom[2][6cm]{\nabstab{-#1}{/\!/}\hspace{0.5em}\textit{#2}} An expressive language with regions will have to support region polymorphism, which invariably leads to \emph{region aliasing}. This must be handled with caution, as a na\"{i}ve approach may cause unsoundness. In the examples that follow, we discuss how region aliasing is used in our language as well as the restrictions that we impose to guarantee safety. Function $\cfont{swap}$ accepts two integer references, residing in regions $\rho_1$ and $\rho_2$, and swaps their contents. It assumes that both regions are already locked and remain locked when the function returns. \begin{ndisplay} \cfont{def} \ \cfont{swap} = \Lambda \nemphthree{\rho_1}. \ \Lambda \nemphthree{\rho_2}. \ \lambda (x : \tref{\nemphthree{\rho_1}}{\cfont{int}}, \ y : \tref{\nemphthree{\rho_2}}{\cfont{int}}). \excom{$\rho_1$ and $\rho_2$ must be both locked} \\ \ntab[1] \cfont{let} \ z = \cfont{deref} \ x \ \cfont{in} \excom{OK: $\rho_1$ is locked} \\ \ntab[2] x := \cfont{deref} \ y; \excom{OK: $\rho_1$ and $\rho_2$ are locked} \\ \ntab[2] y := z \excom{OK: $\rho_2$ is locked} \end{ndisplay} In order to instantiate $\rho_1$ and $\rho_2$ with the same region $\rho$, we can create two lock capabilities by using the $\cfont{lock}$ operator twice on $\rho$'s handle $h$. Of course, the second use of $\cfont{lock}$ will succeed immediately, as the region has already been locked by the same thread. \begin{ndisplay} \ldots \excom{$\{\conelt{\rho}{2,0}{\rho_H}\}$} \\ \cfont{lock} \ \nemphtwo{h}; \ \cfont{lock} \ \nemphtwo{h}; \excom{$\{\conelt{\rho}{2,2}{\rho_H}\}$} \\ \cfont{swap}[\nemphthree{\rho}][\nemphthree{\rho}](a, b); \excom{each $\rho$ parameter requires $\conelt{\rho}{1,1}{\rho_H}$} \\ \cfont{unlock} \ \nemphtwo{h}; \ \cfont{unlock} \ \nemphtwo{h} \excom{$\{\conelt{\rho}{2,0}{\rho_H}\}$} \end{ndisplay} \end{example} \begin{example}[Reentrant locks]\normalfont \newcommand\excom[2][6cm]{\nabstab{-#1}{/\!/}\hspace{0.5em}\textit{#2}} Region aliasing introduces the need for reentrant locks. To see this, let us change the swapping function of the previous example, so that it receives two references in unlocked regions. For swapping their contents, it will have to acquire locks for the two regions (and release them, when they are no longer needed). \begin{ndisplay} \cfont{def} \ \cfont{swap} = \Lambda \nemphthree{\rho_1}. \ \Lambda \nemphthree{\rho_2}.\nbox{% \ \lambda (\nbox{% \nemphtwo{h_1} : \trgn{\nemphthree{\rho_1}}, \ \nemphtwo{h_2} : \trgn{\nemphthree{\rho_2}}. \\ x : \tref{\nemphthree{\rho_1}}{\cfont{int}}, \ y : \tref{\nemphthree{\rho_2}}{\cfont{int}}). } \excom{$\rho_1$ and $\rho_2$ are unlocked}} \\ \ntab[1] \cfont{lock} \ \nemphtwo{h_1}; \\ \ntab[1] \cfont{let} \ z = \cfont{deref} \ x \ \cfont{in} \excom{OK: $\rho_1$ is locked} \\ \ntab[2] \cfont{lock} \ \nemphtwo{h_2}; \\ \ntab[2] x := \cfont{deref} \ y; \excom{OK: $\rho_1$ and $\rho_2$ are locked} \\ \ntab[2] \cfont{unlock} \ \nemphtwo{h_1}; \\ \ntab[2] y := z; \excom{OK: $\rho_2$ is locked} \\ \ntab[2] \cfont{unlock} \ \nemphtwo{h_2} \excom{all locks can be released} \end{ndisplay} Suppose again that we are to instantiate $\rho_1$ and $\rho_2$ with the same region $\rho$. \begin{ndisplay} \ldots \excom{$\{\conelt{\rho}{2,0}{\rho_H}\}$} \\ \cfont{swap}[\nemphthree{\rho}][\nemphthree{\rho}](h, h, a, b); \excom{each $\rho$ parameter requires $\conelt{\rho}{1,0}{\rho_H}$} \end{ndisplay} We can easily see, however, that the run-time system cannot use binary locks; in that case, $\cfont{swap}[\rho][\rho]$ would either come to a deadlock, waiting to obtain once more the lock that it has already acquired, or --- worse --- it might release the lock early (at $\cfont{unlock} \ \nemphtwo{h_1}$) and allow a data race to occur. To avoid unsoundness, we use \emph{reentrant locks}: lock counts are important both for static typing and for the run-time system. A lock with a positive run-time count can immediately be acquired again, if it was held by the same thread. Moreover, a lock is released only when its run-time count becomes zero. \end{example} \begin{example}[Pure and Impure Capabilities]\normalfont \newcommand\excom[2][6cm]{\nabstab{-#1}{/\!/}\hspace{0.5em}\textit{#2}} Unrestricted region aliasing leads to unsoundness. Consider function $\cfont{bad}$, which accepts two integer references ($x$ and $y$) in regions $\rho_1$ and $\rho_2$, which are both locked. It lets $\rho_1$ migrate to a new thread and passes $x$ as a parameter. It then assigns a value to $y$. \begin{ndisplay} \cfont{def} \ \cfont{bad} = \Lambda \nemphthree{\rho_1}. \ \Lambda \nemphthree{\rho_2}. \ \lambda (x : \tref{\nemphthree{\rho_1}}{\cfont{int}}, \ y : \tref{\nemphthree{\rho_2}}{\cfont{int}}). \excom{$\rho_1$ and $\rho_2$ must be both locked} \\ \ntab[1] \cfont{spawn} \ \cfont{f}[\nemphthree{\rho_1}](x); \excom{$\rho_1$ migrates to $\cfont{f}$ while locked} \\ \ntab[1] y := 7 \excom{OK: $\rho_2$ is still locked --- \textrm{WRONG!}} \end{ndisplay} A data race may occur if we call $\cfont{bad}$ as follows; both threads have access to $a$, each holding a lock for $\rho$. \begin{ndisplay} \cfont{swap}[\nemphthree{\rho}][\nemphthree{\rho}](a, a); \excom{each $\rho$ parameter requires $\conelt{\rho}{1,1}{\rho_H}$} \end{ndisplay} The cause of the unsoundness is that, in this last call to $\cfont{swap}[\rho][\rho]$, we allowed a single capability $\conelt{\rho}{2,2}{\rho_H}$ to be divided in two distinct capabilities $\conelt{\rho}{1,1}{\rho_H}$. More specifically, we divided the lock count in two and created two distinct lock capabilities, one of which escaped to a different thread through region migration. To resolve the unsoundness, we introduce the notion of $\emph{pure}$ (i.e., full) and $\emph{impure}$ (i.e., divided) capabilities. For instance, $\conelt{\rho}{2,2}{\rho_H}$ is a pure capability; when we divide it we obtain two impure halves, which we denote as $\conelt{\rho}{\overline{1,1}}{\rho_H}$. Impure capabilities cannot be given to newly spawned threads when their lock count is positive. In contrast with pure capabilities, they represent inexact knowledge of a region's counts. \end{example} \section{Language Description} \label{sec:language} The syntax of the language is illustrated in Figure~\ref{fig:syntax:core}. The language core comprises of variables ($x$), constants ($c$), functions, and function application. Functions can be region polymorphic ($\opoly{\rho}{f}$) and region application is explicit ($e[\rho]$). Monomorphic functions ($\lambda x.\, e$) must be annotated with their type. The application of monomorphic functions is annotated with a \emph{calling mode} ($\xi$), which is $\nL$ for normal (sequential) application and $\nP{\gamma}$ for spawning a new thread.% \footnote{In the examples of Section~\ref{sec:examples}, we used more intuitive notation: we omitted $\nL$ and used the keyword $\cfont{spawn}$ instead of $\mathsf{par}$.} Parallel application is annotated with the input effect of the new thread ($\gamma$); this annotation can be automatically inferred by the type checker. The constructs for manipulating references are standard. A newly allocated memory cell is returned by $\onew{e_1}{e_2}{}$, where $e_1$ is the value that will be placed in the cell and $e_2$ is a handle of the region in which the new cell will be allocated. Standard assignment and dereference operators complete the picture. \begin{figure} \begin{minipage}[t]{10cm} \gbox{1.75cm}{2cm} { \grFunc \grExpr \grType \grConstraint } \end{minipage} \hfill \begin{minipage}[t]{5.5cm} \gbox{2.5cm}{2cm} { \grScope \grCapDelta \grCapKind \grCap \grParent } \end{minipage} \caption{Syntax.\label{fig:syntax:core}} \end{figure}% The construct $\onewrgn{\rho}{x}{e_1}{e_2}$ allocates a new region $\rho$ and binds $x$ to the region \emph{handle}. The new region resides in a \emph{parent} region, whose handle is given in $e_1$. The scope of $\rho$ and $x$ is $e_2$, which must consume the new region by the end of its execution. A region can be consumed either by deallocation or by transferring its ownership to another thread. At any given program point, each region is associated with a \emph{capability} ($\kappa$). Capabilities consist of two natural numbers, the \emph{capability counts}: the \emph{region} count and \emph{lock} count, which denote whether a region is live and locked respectively. When first allocated, a region starts with capability $(1, 1)$, meaning that it is live and locked, so that it can be exclusively accessed by the thread that allocated it. As we have seen, this is our equivalent of a thread-local region. By using the construct $\ocap{\psi}{\eta}{e}$, a thread can \emph{increment} or \emph{decrement} the capability counts of the region whose handle is specified in $e$. The capability operator $\eta$ can be, e.g., $\nRg+$ (meaning that the region count is to be incremented) or $\nLk-$ (meaning that the lock count is to be decremented).% \footnote{The region manipulation operators used in Section~\ref{sec:examples} are simple abbreviations: $\cfont{share} \equiv \cfont{cap}_{\nRg+}$, $\cfont{unlock} \equiv \cfont{cap}_{\nLk-}$, etc. } When a region count reaches zero, the region may be physically deallocated and no subsequent operations can be performed on it. When a lock count reaches zero, the region is unlocked. As we explained, capability counts determine the validity of operations on regions and references. All memory-related operations require that the involved regions are live, i.e., the region count is greater than zero. Assignment and dereference can be performed only when the corresponding region is live and locked. A capability of the form $(n_1, n_2)$ is called a \emph{pure} capability, whereas a capability of the form $(\overline{n_1,n_2})$ is called an \emph{impure} capability. In both cases, it is implied that the current thread can decrement the region count $n_1$ times and the lock count $n_2$ times. Impure capabilities are obtained by splitting pure or other impure capabilities into several pieces, e.g., $(3, 2) = (\overline{2, 1}) + (\overline{1, 1})$, in the same spirit as \emph{fractional capabilities}~\cite{Boyland@SAS-03}. As we explained in Example~9 of Section~\ref{sec:examples}, these pieces are useful for region aliasing, when the same region is to be passed to a function in the place of two distinct region parameters. An impure capability implies that our knowledge of the region and lock count is inexact. The use of such capabilities must be restricted; e.g., an impure capability with a non-zero lock count cannot be passed to another thread, as it is unsound to allow two threads to simultaneously hold the same lock. Capability splitting takes place automatically with function application. \section{Operational Semantics} \label{sec:operational} We define a \emph{small-step} operational semantics for our language, using two evaluation relations, at the level of \emph{threads} and \emph{expressions} (Figures~\ref{fig:threadeval} and \ref{fig:expeval}% \if\thepage=\pageref{fig:expeval}\else \ on the next page% \fi). The thread evaluation relation transforms \emph{configurations}. A configuration $C$ (see Figure~\ref{fig:eval}) \begin{figure} \begin{minipage}[t]{7cm} \gbox{2.5cm}{2cm}{ \grConf \grThread \grRList \grCounts \grContents } \end{minipage} \hfill \begin{minipage}[t]{7cm} \gbox{2.25cm}{2cm}{ \grEvalCont } \end{minipage} \caption{Configurations, store, threads and evaluation contexts.% \label{fig:eval}} \end{figure}% consists of an abstract \emph{store} $S$ and a list of threads $T$.% \footnote{The order of elements in comma-separated lists, e.g.\ in a store $S$ or in a list of threads $T$, is not important; we consider all list permutations as equivalent.} Each thread in $T$ is of the form $n: e$, where $n$ is a thread identifier and $e$ is an expression. The store is a list of regions of the form $\imath:(\theta,H,S)$, where $\imath$ is a \emph{region identifier}, $\theta$ is a thread map, $H$ is a memory heap and $S$ is the list of subregions in the region hierarchy. The thread map associates thread identifiers with capability counts for region $\imath$, whereas the memory heap represents the region's contents, mapping locations to values. A \emph{thread evaluation context} $E$ (Figure~\ref{fig:eval}) is defined as an expression with a \emph{hole}, represented as $\opbox$. The hole indicates the position where the next reduction step can take place. Our notion of evaluation context imposes a call-by-value evaluation strategy to our language. Subexpressions are evaluated in a left-to-right order. We assume that concurrent reduction events can be totally ordered. At each step, a random thread ($n$) is chosen from the thread list for evaluation (Figure~\ref{fig:threadeval}). \begin{figure*}\centering \tboxx{440pt}{ \orSpawn \\[2em] \orPickOne \ensuremath{\hspace{2em}} \orTerminate } \caption{Thread evaluation relation $\,C \oparrowt{n} C \,'$.% \label{fig:threadeval}} \vspace*{6pt} \end{figure*}% It should be noted that the thread evaluation rules are the only \emph{non-deterministic} rules in the operational semantics of our language; in the presence of more than one active threads, our semantics does not specify which one will be selected for evaluation. Threads that have completed their evaluation and have been reduced to \emph{unit} values, represented as $\ounit$, are removed from the active thread list (rule \nrulelabel{E-T}). Rule~\nrulelabel{E-S} reduces some thread $n$ via the expression evaluation relation. When a parallel function application redex is detected within the evaluation context of a thread, a new thread is created (rule \nrulelabel{E-SN}). The redex is replaced with a unit value in the currently executed thread and a new thread is added to the thread list, with a \emph{fresh} thread identifier. The \emph{partial} function $\mathrm{transfer}(\cprog,n,n',\gamma_1)$ updates the thread maps of all regions specified in $\gamma_1$, transferring capabilities between threads $n$ and $n'$. It is undefined when this transfer is not possible. The expression evaluation relation is defined in Figure~\ref{fig:expeval}. \begin{figure*}\centering \tboxx{532pt}{ \orApp \ensuremath{\hspace{2em}} \orRPoly \\[2em] \orNewReg \ensuremath{\hspace{2em}} \orCapA \\[2em] \orNewRef \ensuremath{\hspace{2em}} \orAsgn \ensuremath{\hspace{2em}} \orDeref } \caption{Expression evaluation relation $\,\cprog;e \oparrowe{n} \, \cprog';e'$.% \label{fig:expeval}} \end{figure*}% The rules for reducing function application (\nrulelabel{E-A}) and region application (\nrulelabel{E-RP}) are standard. The remaining rules make use of five \emph{partial} functions that manipulate the store. These functions are undefined when their constraints are not met. All of them require that some region is \emph{live}. A region is \emph{live} when the sum of all region counts in the thread map associated with that region is positive and all ancestors of the region are \emph{live} as well. In addition to liveness, some of these functions require that some region is \emph{accessible} to the currently executed thread. Region $r$ is accessible to some thread $n$ (and \emph{inaccessible} to all other threads) when $r$ is live and the thread map associated with $r$, or with some ancestor of $r$, maps $n$ to a positive lock count. \begin{itemize} \item $\mathrm{alloc}(\ensuremath{\cfont{j}}, \cprog, v)$ is used in rule \nrulelabel{E-NR} for creating a new reference. It allocates a new object in $\cprog$. The object is placed in region $\ensuremath{\cfont{j}}$ and is set to value $v$. Region $\ensuremath{\cfont{j}}$ must be live. Upon success, the function returns a pair $(\cprog',\ell)$ containing the new store and a fresh location for the new object. \item $\mathrm{lookup}(\cprog,\ell,n)$ is used in rule \nrulelabel{E-D} to look up the value of location $\ell$ in $\cprog$. The region in which $\ell$ resides must be accessible to the currently executed thread $n$. Upon success, the function returns the value $v$ stored at $\ell$. \item $\mathrm{update}(\cprog, \ell,v,n)$ is used in rule \nrulelabel{E-AS} to assign the value $v$ to location $\ell$ in $\cprog$. The region in which $\ell$ resides must be accessible to the currently executed thread $n$. Upon success, the function returns the new store $\cprog'$. \item $\mathrm{newrgn}(\cprog,n,\ensuremath{\cfont{j}})$ is used in rule \nrulelabel{E-NG} to create a new region in $\cprog$. The new region is allocated within $\ensuremath{\cfont{j}}$, which must be live. Its thread map is set to $n \mapsto 1,1$. Upon success, the function returns a pair $(\cprog',k)$ containing the new store and a fresh region name for the new region. \item $\mathrm{updcap}(\cprog,\eta,\ensuremath{\cfont{j}},n)$ is used in rule \nrulelabel{E-C}. This operation updates $\cprog$ by modifying the region or lock count of thread $n$ for region $\ensuremath{\cfont{j}}$. Upon success, the function returns the new store $\cprog'$. When a lock update is requested and the lock is held by another thread, the result is undefined. In this case, rule \nrulelabel{E-C} cannot be applied and the operation will block, until the lock is available. \end{itemize} The operational semantics may get stuck when a \emph{deadlock} occurs. Our semantics will also get stuck when a thread attempts to access a memory location without having acquired an appropriate lock for this location. In this case, $\mathrm{update}(\cprog, \ell,v,n)$ and $\mathrm{lookup}(\cprog, \ell, n)$ are undefined and it is impossible to perform a single step via rules \nrulelabel{E-AS} or \nrulelabel{E-D}. The same is true in several other situations (e.g.\ when referring to a non-existent region or location). Threads that may cause a data race will definitely get stuck. We follow a different approach from related work, e.g.\ the work of Grossman \cite{Grossman@TLDI-03}, where a special kind of value $\mathit{junk}_v$ is often used as an intermediate step when assigning a value $v$ to a location, before the real assignment takes place, and type safety guarantees that no junk values are ever used. As described above, we use a more direct approach by incorporating the locking mechanism in the operational semantics. Our progress lemma in Section~\ref{sec:safety} guarantees that, at any time, \emph{all} threads can make progress and, therefore, a possible implementation does not need to check liveness or accessibility at run-time. \section{Static Semantics} \label{sec:typing} In this section we discuss the most interesting parts of our type system. As we sketched in Section~\ref{sec:examples}, to enforce our safety invariants, we use a \emph{type and effect system}. Effects are used to statically track region capabilities. An effect ($\gamma$) is a list of elements of the form $\conelt{r}{\kappa}{\pi}$, denoting that region $r$ is associated with capability $\kappa$ and has parent $\pi$, which can be another region, $\bot$, or $\mathrm{?}$. Regions whose parents are $\bot$ or $\mathrm{?}$ are considered as roots in our region hierarchy. We assume that there is an initial (physical) root region corresponding to the entire heap, whose handle is available to the main program. The parent of the heap region is $\bot$. More (logical) root regions can be created using hierarchy abstraction. The abstract parent of a region that is passed to a function is denoted by $\mathrm{?}$. \renewcommand\mathrm[1]{\ensuremath{\text{\textsl{#1}}}} \begin{figure*} \centering% \tboxx{\textwidth}{% \trFunc \ensuremath{\hspace{2em}} \trApp \\[2em] \trDeref \ensuremath{\hspace{2em}} \trNewRgn \\[2em] \trNewRef \ensuremath{\hspace{2em}} \trCap } \caption{Selected typing rules.\label{fig:typrules}} \end{figure*}% \begin{figure*} \centering% \tboxx{\textwidth}{% \trEffSplitJoin \\[2em] \trCapSplitJoin } \caption{Effect and capability splitting.\label{fig:splitjoinrules}} \end{figure*}% The syntax of types in Figure~\ref{fig:syntax:core} (on page \pageref{fig:syntax:core}) is more or less standard. A collection of base types $b$ is assumed; the syntax of values belonging to these types and operations upon such values are omitted from this paper. We assume the existence of a \emph{unit} base type, which we denote by $\tunit$. Region handle types $\trgn{r}$ and reference types $\tref{\tau}{r}$ are associated with a type-level region $r$. Monomorphic function types carry an \emph{input} and an \emph{output effect}. A well-typed expression $e$ has a type $\tau$ under an input effect $\gamma$ and results in an output effect $\gamma'$. The typing relation (see~Figure~\ref{fig:typrules}) is denoted by $\ttype{\cstdall}{e}{\tau}{\gamma}{\gamma'}$ and uses four typing contexts: a set of region literals ($R$), a mapping of locations to types ($M$), a set of region variables ($\Delta$), and a mapping of term variables to types ($\Gamma$). The effects that appear in our typing relation must satisfy a \emph{liveness invariant}: all regions that appear in the effect are \emph{live}, i.e., their region counts and those of all their ancestors are positive. Thus, in order to check if a region $r$ is live in the effect $\gamma$, we only need to check that $r \in \mathrm{dom}(\gamma)$. The typing rule for lambda abstraction (\nrulelabel{T-F}) requires that the body $e$ is well-typed with respect to the effects ascribed on its type. The typing rule for function application (\nrulelabel{T-AP}) splits the output effect of $e_2$ ($\gamma''$) by subtracting the function's input effect ($\gamma_1$). It then joins the remaining effect with the function's output effect ($\gamma_2$). In the case of parallel application, rule \nrulelabel{T-AP} also requires that the return type is unit. The splitting and joining of effects is controlled by the judgement $\effeq{\xi}{\gamma'' = \gamma_2 \ensuremath{\oplus} (\gamma \ensuremath{\ominus} \gamma_1)}$, which is defined in Figure~\ref{fig:splitjoinrules} (the auxiliary functions and predicates are defined in Figures~\ref{fig:auxpred} and~\ref{fig:auxiliary}). \begin{figure*}[t] \centering% \tboxx{\textwidth}{% \trLive \\[2em] \trAccessible } \caption{Auxiliary predicates: region liveness and accessibility.% \label{fig:auxpred}} \end{figure*}% \begin{figure*}[t] \centering% \newcommand\nset[1]{\ensuremath{\left\{\, #1 \,\right\}}} \newcommand\nwhere{\ensuremath{\ \mid\ }} \tboxx{\textwidth} \trDefs } \caption{Auxiliary functions and predicates.\label{fig:auxiliary}} \end{figure*}% It enforces the following properties: \bgroup\noitemsep \begin{itemize} \item the liveness invariant for $\gamma''$; \item the consistency of $\gamma$ and $\gamma''$, i.e., regions cannot change parent and capabilities cannot switch from pure to impure or vice versa; the domain of $\gamma''$ is a subset of the domain of $\gamma$; \item for sequential application, all parent regions that become abstracted for the duration of the function call must be live after the function returns; \item for parallel application, the thread output effect must be empty, the thread input effect must not contain impure capabilities with positive lock counts and hierarchy abstraction is disallowed. \end{itemize} \egroup The typing rules for references are standard. In Figure~\ref{fig:typrules} we only show the rules for dereference (\nrulelabel{T-D}) and reference allocation (\nrulelabel{T-NR}). The former checks that region $r$ is \emph{accessible}. The latter only checks that the region $r$ is live. The rule for creating new regions (\nrulelabel{T-NG}) checks that $e_1$ is a handle for some live region $r'$. Expression $e_2$ is type checked in an extended typing context (i.e., $\rho$ and $x:\trgn{\rho}$ are appended to $\Delta$ and $\Gamma$ respectively) and an extended input effect (i.e., a new effect is appended to the input effect such that the new region is live and accessible to this thread). The rule also checks that the type and the output effect of $e_2$ do not contain any occurrence of region variable $\rho$. This implies that $\rho$ must be \emph{consumed} by the end of the scope of $e_2$. The capability manipulation rule (\nrulelabel{T-CP}) checks that $e$ is a handle of a live region $r$. It then modifies the capability count of $r$ as dictated by function $\[[\eta\]]$, which increases or decreases the region or the lock count of its argument, according to the value of $\eta$. The dynamic semantics ensures that an operational step is performed when the actual counts are consistent with the desired output effect. For instance, if the lock of region $r$ is held by some other executing thread, the evaluation of $\ocap{}{\nLk+}{}$\relax must be suspended until the lock can be obtained. On the other hand, the evaluation of $\ocap{}{\nRg-}{}$\relax does not need to suspend but may not be able to physically deallocate a region, as it may be used by other threads. \section{Type Safety} \label{sec:safety} In this section we discuss the fundamental theorems that prove type safety of our language.% \footnote{% Full proofs and a full formalization of our language are given \iftechrep in the Appendix. \else in the companion technical report \cite{ReglockTypeSoundness}. \fi } The type safety formulation is based on proving the \emph{preservation} and \emph{progress} lemmata. Informally, a program written in our language is safe when for each thread of execution an evaluation step can be performed or that thread is waiting for a lock (\emph{blocked}). As discussed in Section~\ref{sec:operational}, a thread may become stuck when it accesses a region that is not live or accessible (these are obviously the interesting cases in our concurrent setting; of course a thread may become stuck when it performs a non well-typed operation). Deadlocked threads are not considered to be stuck. \begin{ndefnn}[Thread Typing] Let $T$ be a collection of threads. Let $R;M;\delta$ be a global typing context, in which $\delta$ is a mapping from thread identifiers to effects, used only for metatheoretic purposes. For each thread $\oth{n}{e}$ in $T$, we take $\rfl{\delta}{n}$ to be the input effect that corresponds to the evaluation of expression $e$. The following rules define \emph{well-typed} threads. \begin{nruledisplay} \trThreads \end{nruledisplay} \end{ndefnn} \begin{ndefnn}[Store Consistency] A store $S$ is \emph{consistent} with respect to an effect mapping $\delta$ when the following conditions are met: \begin{itemize} \item \emph{Region consistency}: the set of region names occurring in the co-domain of $\delta$ is a subset of the set of region names in $S$. \item \emph{Static-dynamic count consistency}: for each region, the dynamic region and lock counts of some thread must be greater than or equal to the corresponding static counts of the same thread. \item \emph{Mutual exclusion}: only one thread may have a positive lock count in $\delta$ for a particular region $\ensuremath{\cfont{j}}$. Additionally, only this thread is allowed to access or lock sub-regions of $\ensuremath{\cfont{j}}$. \end{itemize} \end{ndefnn} \begin{ndefnn}[Store Typing] A store $S$ is \emph{well-typed} with respect to $R;M;\delta$ (we denote this by\linebreak \rstr{M}{R}{\delta}{P}{S}) when the following conditions are met: \begin{itemize} \item $S$ is \emph{consistent} with respect to $\delta$, \item the set of region names in $S$ is equal to $R$, \item the set of locations in $M$ is equal to the set of locations in $S$, and \item for each location $\ell$, the stored value $S(\ell)$ is closed and has type $M(\ell)$ with empty effects, i.e., \ttype{R;M;\emptyset;\emptyset}{S(\ell)}{M(\ell)}{\emptyset}{\emptyset}. \end{itemize} \end{ndefnn} \begin{ndefnn}[Configuration Typing] A configuration $S;T$ is \emph{well-typed} with respect to $R;M;\delta$ (we denote this by $\tctype{R;M;\delta}{S;T}$) when the collection of threads $T$ is well-typed with respect to $R;M;\delta$ and the store $S$ is well-typed with respect to $R;M;\delta$. % \end{ndefnn} \begin{ndefnn}[Not stuck] A configuration $S;T$ is \emph{not stuck} when each thread in $T$ can take one of the evaluation steps in Figure~\ref{fig:threadeval} (\nrulelabel{E-S}, \nrulelabel{E-T} or \nrulelabel{E-SN}) or it is waiting for a lock held by some other thread. \end{ndefnn} Given these definitions, we can now present the main results of this paper. The \emph{progress} and \emph{preservation} lemmata are first formalized at the \emph{program} level, i.e., for all concurrently executed threads. \begin{lemma}[Progress --- Program]% \label{lemma:thread_progress}\normalfont Let $S;T$ be a closed well-typed configuration with $\tctype{R;M;\delta}{S;T}$. Then $S;T$ is not stuck. \end{lemma} \begin{lemma}[Preservation --- Program]% \label{lemma:thread_preservation}\normalfont Let $S;T$ be a well-typed configuration with $\tctype{R;M;\delta}{S;T}$. If the operational semantics takes a step $S;T \oparrowt{\imath} S';T'$, then there exist $R' \supseteq R$, $M' \supseteq M$ and $\delta'$ such that the resulting configuration is well-typed with $\tctype{R';M';\delta'}{S';T'}$. \end{lemma} An expression-level version for each of these two lemmata is required, in order to prove the above. At the \emph{expression} level, progress and preservation are defined as follows. \begin{lemma}[Progress --- Expression]% \label{lemma:thread_progress_local}\normalfont Let $S$ be a well-typed store with $\rstr{M}{R}{\delta,n \mapsto \gamma} {P}{S}$ and let $e$ be a closed well-typed \emph{redex} with $ \ttype{R;M;\emptyset;\emptyset}{e}{\tau} {\gamma}{\gamma'} $. Then exactly one of the following is true: \begin{itemize} \item $e$ is of the form $\ocap{\nLk}{\nLk+}{\orgn{\ensuremath{\cfont{j}}}}$ and $\ensuremath{\cfont{j}}$ is a live but inaccessible region to thread $n$, or \item $e$ is of the form $\oapp{ \ofunc{x}{e_1}{\tau}}{ v}{\nP{\gamma}}$ or \item there exist $S'$ and $e'$ such that $S;e \, \oparrowe{n} S';e'$. \end{itemize} \end{lemma} \begin{lemma}[Preservation --- Expression]% \label{lemma:thread_preservation_local}\normalfont Let $e$ be a well-typed expression with $\ttype{R;M;\emptyset;\emptyset}{e}{\tau}{\gamma}{\gamma''}$ and let $S$ be a well-typed store with $\rstr{M}{R}{\delta,n \mapsto \gamma}{}{S}$. If the operational semantics takes a step $S;e \oparrowe{n} S';e'$, then there exist $R' \supseteq R$, $M' \supseteq M$ and $\gamma'$ such that the resulting expression and the resulting store are well-typed with $\ttype{R';M';\emptyset;\emptyset}{e'}{\tau}{\gamma'}{\gamma''}$ and $\rstr{M'}{R'}{\ops{\delta}{n}{\gamma'}}{}{S'}$. \end{lemma} The \emph{type safety} theorem is a direct consequence of Lemmata~\ref{lemma:thread_progress} and~\ref{lemma:thread_preservation}. Let function $\cfont{main}$ be the initial program, let $\iota_H$ be global heap region and let the initial typing contexts $R_0$ and $\delta_0$ and the initial program configuration $S_0; T_0$ be defined by the following singleton lists: \iftechrep \begin{ndisplay}[max={R_0}] \ndefine{R_0}{=}{% \{ \iota_H \} } \\ \ndefine{\delta_0}{=}{% \{ 1 \mapsto \conelt{\iota_H}{1,0}{\bot} \} } \\ \ndefine{\theta_0}{=}{% \{ 1 \mapsto 1,0 \} } \\ \ndefine{S_0}{=}{% \{ \iota_H:(\theta_0, \emptyset, \emptyset) \} } \\ \ndefine{T_0}{=}{% \{ 1 : \oapp{\cfont{main}[\iota_H]}{\orgn{\iota_H}}{\nL{}} \} } \end{ndisplay} \else $R_0 = \{ \iota_H \}$, $\delta_0 = \{ 1 \mapsto \conelt{\iota_H}{1,0}{\bot} \}$, $\theta_0 = \{ 1 \mapsto 1,0 \}$, $S_0 = \{ \iota_H:(\theta_0, \emptyset, \emptyset) \}$, and $T_0 = \{ 1 : \oapp{\cfont{main}[\iota_H]}{\orgn{\iota_H}}{\nL{}} \}$. \fi \begin{theorem}[Type Safety]% \label{theorem:type_safety}\normalfont If the initial configuration $S_0;T_0$ is well-typed with $\tctype{R_0;\emptyset;\delta_0}{S_0;T_0}$ and the operational semantics takes any number of steps $S_0;T_0 \oparrowt{n}^{n} S_n;T_n$, then the resulting configuration $S_n;T_n$ is not stuck. \end{theorem} The empty (except for $R_0$ that contains only $\iota_H$) contexts that are used when typechecking the initial configuration $S_0;T_0$ guarantee that all functions in the program are closed and that no explicit region values ($\orgn{\imath}$) or location values ($\oloc{\ell}$) are used in the source of the original program. \section{Related Work} \label{sec:related} The first statically checked stack-based region system was developed by Tofte and Talpin~\cite{TofteTalpin@POPL-94}. Since then, several memory-safe systems that enabled early region deallocation for a sequential language were proposed~\cite{AikenFahndrichLevien@PLDI-95,% RegionDeallocHenglein@PPDP-01,% RegionsLinearTypes@ICFP-01,% LinearRegions@ESOP-06% }. Cyclone~\cite{Cyclone@PLDI-02} and RC~\cite{GayAiken@PLDI-01} were the first imperative languages to allow safe region-based management with explicit constructs. Both allowed early region deallocation and RC also introduced the notion of multi-level region hierarchies. RC programs may throw region-related exceptions, whereas our approach is purely static. Both Cyclone and RC make no claims of memory safety or race freedom for concurrent programs. Grossman proposed a type system for safe multi-threading in Cyclone~\cite{Grossman@TLDI-03}. Race freedom is guaranteed by statically tracking locksets within lexically-scoped synchronization constructs. Grossman's proposal allows for fine-grained locking, but only deals with stack-based regions and does not enable early release of regions and locks. In contrast, we support hierarchical locking, as opposed to just primitive locking, and bulk region deallocation. Statically checked region systems have also been proposed~\cite{OwnershipTypes@PLDI-03,% ScopedTypes@RTSS-04,% ImplicitOwnershipTypes@SCP-08% } for real-time Java to rule out dynamic checks imposed by the language specification. Boyapati et al.~\cite{OwnershipTypes@PLDI-03} introduce hierarchical regions in ownership types but the approach suffers from the same disadvantages as Grossman's work. Additionally, their type system only allows sub-regions for \emph{shared} regions, whereas we do not have this limitation. Boyapati also proposed an ownership-based type system that prevents deadlocks and data races~\cite{OwnershipTypes@OOPSLA-02}; in contrast to his system, we support locking of arbitrary nodes in the region hierarchy. Static region hierarchies (depth-wise) have been used by Zhao~\cite{ScopedTypes@RTSS-04}. Their main advantage is that programs require fewer annotations compared to programs with explicit region constructs. In the same track, Zhao et al.~\cite{ImplicitOwnershipTypes@SCP-08} proposed implicit ownership annotations for regions. Thus, classes that have no explicit owner can be allocated in any static region. This is a form of \emph{existential ownership}. In contrast, we allow a region to completely abstract its owner/ancestor information by using the \emph{hierarchy abstraction} mechanism. None of the above approaches allow full ownership abstraction for region subtrees. Cunningham et al.~\cite{UniverseRaces@VAMP-07} proposed a universe type system to guarantee race freedom in a calculus of objects. Similarly to our system, object hierachies can be atomically locked at any level. Unlike our system, they do not support early lock releases and lock ownership transfers between threads. Consequently, their system cannot encode two important aspects of multi-threaded programming: thread-locality and data migration. Finally, our system provides explicit memory management and supports separate compilation. The main limitation of our work is that we require explicit annotations regarding ownership and region capabilities. Moreover, our locking system offers coarser-grained locking than most other related works. The use of hierarchical locking avoids some, though not all, deadlocks. \section{Concluding Remarks} \label{sec:concluding} In this paper, we have presented a concurrent language emloying region-based memory management and locking primitives. Regions and locks are organized in a common hierarchy and treated uniformly. Our language allows atomic deallocation and locking of entire subtrees at any level of the hierarchy; it also allows region and lock capabilities to be transferred between threads, encoding useful idioms of concurrent programming such as \emph{thread-local data} and \emph{data migration}. The type system guarantees the absence of memory access violations and data races in the presence of region aliasing. We are currently integrating our system in Cyclone. In the future, we are planning to extend our type system to achieve an exact cor\-re\-spon\-dence between static and dynamic capability counts, and provide deadlock freedom guarantees.
\section{Introduction} After the observations at the Relativistic Heavy Ion Collider (RHIC) that the elliptic flow measurements are consistent with ideal fluid dynamics predictions \cite{Kolb:2000fha,Huovinen:2001cy,Csernai:2003xd} the development of numerical solutions of viscous hydrodynamics equations is actively pursued \cite{Song:2007fn,Song:2007ux,Luzum:2008cw,Luzum:2009sb,Song:2009rh}. The idea of those calculations is to quantify the deviations from local thermal equilibrium by comparisons to the available data for spectra and elliptic flow. Solving the differential equations of any kind of fluid dynamics prescription implies the knowledge of the boundary conditions, i.e. the initial conditions and the freeze-out criterion. Therefore, it is essential to investigate in a systematic way how different initial profiles and freeze-out implementations influence the final observable results. This article focuses on this question concerning the initial conditions. The realistic situation of the collision of two nuclei suffers from many different sources of initial state fluctuations \cite{Hirano:2009ah}. The density profiles are not smooth, but there are peaks in the transverse and the longitudinal direction. There are impact parameter fluctuations within one specific centrality class leading to multiplicity fluctuations and differences in the initial geometry \cite{Broniowski:2009fm}. Furthermore, the nuclei do not necessarily collide in the event-plane given by the laboratory system, but might also have a rotated reaction plane. All these effects are averaged out, if assuming a smooth symmetric initial density profile as it is widely done by parametrising the initial conditions for hydrodynamic calculations within a Glauber or Color Glass Condensate (CGC) picture. To avoid potential misunderstandings, Monte-Carlo Glauber approaches (and CGC approaches) that produce fluctuating initial conditions are available and widely used, e.g. by experimental collaborations. However, usually these fluctuating initial conditions are not used as event-by-event initial conditions for the majority of hydrodynamic simulations. For alternative hydrodynamic approaches with fluctuating initial conditions see e.g. \cite{Andrade:2006yh,Werner:2009fa}. In this article an integrated Boltzmann+hydrodynamics transport approach is applied to the simulation of heavy ion reactions in the energy regime from $E_{\rm lab}=2-160A$ GeV. The initial conditions are generated by the Ultra-relativistic Quantum Molecular Dynamics (UrQMD) approach \cite{Bass:1998ca,Bleicher:1999xi} and the above mentioned event-by-event fluctuations are further propagated in the ideal hydrodynamic evolution employed for the hot and dense stage of the collision. On the other hand, calculations with averaged initial conditions are performed using the very same general setup. Comparing these calculations to the fluctuating setup the effect on the final observable elliptic flow is estimated. \section{Initial Eccentricity} Let us start by looking at the initial state eccentricity as it is given by the UrQMD approach at the starting time $t_{\rm start} = {2R}/{\sqrt{\gamma^2 -1}}$ of the hydrodynamic evolution. For the specific values of these times for Au+Au/Pb+Pb collisions at different beam energies see Table \ref{tab_tstart}. This is the earliest possible transition time at which local equilibrium might have been established after the two nuclei have passed through each other. \begin{table}[h] \begin{center} \renewcommand{\arraystretch}{1.3} \begin{tabular}{|c|c|} \hline $E_{\rm lab}$ [GeV/nucleon] & $t_{\rm start}$ [fm]\\ \hline \hline 2 & 12.417\\ 6 & 7.169 \\ 11 & 5.295 \\ 40 & 2.830\\ 160 & 1.415\\ \hline \end{tabular} \caption{\label{tab_tstart} Starting times of the hydrodynamic evolution for Au+Au/Pb+Pb collisions at different beam energies. The eccentricities displayed in Fig. \ref{ini_ecc_exc} are also calculated at these times.} \end{center} \end{table} The eccentricity quantifies the spatial anisotropy of the initial state which is transformed via pressure gradients into a momentum space anisotropy in the transverse plane that can be quantified by the elliptic flow coefficient. The standard definition for the eccentricity is the reaction plane eccentricity \begin{equation} \epsilon_{\rm RP} = \frac{\sigma_y^2-\sigma_x^2}{\sigma_y^2+\sigma_x^2} \quad, \end{equation} with $\sigma_x^2=\langle x^2 \rangle -\langle x \rangle^2$ and $\sigma_y^2=\langle y^2 \rangle - \langle y \rangle^2$. Especially for smaller colliding systems the so called participant eccentricity is popular \begin{equation} \epsilon_{\rm part}=\frac{\sqrt{(\sigma_y^2-\sigma_x^2)^2+4 \sigma_{xy}^2}}{\sigma_y^2+\sigma_x^2} \quad , \end{equation} where an additional correlation term $\sigma_{xy}=\langle xy \rangle - \langle x \rangle \langle y \rangle $ is introduced. The participant eccentricity incorporates the fact that the participants themselves might be rotated with respect to the reaction plane. The averages $\langle \cdot \rangle$ indicate averages over all particles in one event and the mean eccentricity is obtained by averages over many UrQMD events. The standard deviation (shown as error bars in Fig.\ref{ini_ecc_exc}) with respect to the event average is $\delta \epsilon=\langle \epsilon^2 \rangle -\langle \epsilon \rangle^2$. \begin{figure}[b] \vspace{-1cm} \resizebox{0.5\textwidth}{!}{ \centering \includegraphics{ini_ecc_fluc_exc.eps} } \caption{(Color online) Initial eccentricity $\epsilon_{\rm RP}$ (full lines) and $\epsilon_{\rm part}$ (dotted lines) at time $t_{\rm start}$ for Au+Au/Pb+Pb collisions as a function of the beam energy. The results are shown for four different centrality selections, $b < 3.4$ fm, $b=5-9$ fm, $b=9-11$ fm and fixed impact parameter $b=7$ fm. The standard deviation $\delta \epsilon$ is depicted as error bars to the mean value.} \label{ini_ecc_exc} \end{figure} Figs. \ref{ini_ecc_exc} and \ref{ini_ecc_imp} show the initial eccentricity in dependence of the beam energy for different centralities and the impact parameter dependence for two different beam energies, respectively. The full lines depict the reaction plane eccentricity while the dotted lines refer to the participant eccentricity. In these calculations all particles which have suffered at least one interaction are included. This is consistent with the particles that are taken into account for the hydrodynamic evolution where the spectators are propagated separately in the cascade. The initial eccentricity - mean value and standard deviation (as error bars) - as it is shown in Fig. \ref{ini_ecc_exc} increases with the beam energy for both definitions. For simplicity the starting times are held constant with respect to the centrality variation. As expected the eccentricity is larger for more peripheral collisions. The participant eccentricity is always larger than the reaction plane eccentricity since it is calculated in the frame where the eccentricity is the largest possible one. Also the fluctuations of the eccentricity grow with increasing beam energy and with decreasing centrality of the collision. The fixed impact parameter calculation for $b=7$ fm leads to very similar results as the $b=5-9$ fm calculation for mid-central events. This hints to the fact, that centrality fluctuations have only a minor impact on the overall event-by-event fluctuations of the initial state. \begin{figure}[t] \resizebox{0.5\textwidth}{!}{ \centering \includegraphics{ini_ecc_fluc_imp_sps.eps} } \caption{(Color online) Initial eccentricity $\epsilon_{\rm RP}$ (full lines) and $\epsilon_{\rm part}$ (dotted lines) at time $t_{\rm start}$ for Pb+Pb collisions as a function of the impact parameter for $E_{\rm lab}=40A$ GeV (a) and $E_{\rm lab}=160A$ GeV (b). The standard deviation $\delta \epsilon$ is depicted as error bars to the mean value.} \label{ini_ecc_imp} \end{figure} Apart from the fact that the initial eccentricity at the highest SPS energy (see Fig. \ref{ini_ecc_imp}) is larger than at $E_{\rm lab}=40A$ GeV the results for both energies are pretty similar. The reaction plane eccentricity is zero for small impact parameters while the participant eccentricity stays finite due to the rotation to the participant plane. Going to peripheral collisions the participant eccentricity increases almost linearly and the reaction plane eccentricity drops down again. The fluctuations grow much more for the reaction plane definition beyond an impact parameter of $b=10$ fm. The most interesting observation from Fig. \ref{ini_ecc_imp} is the small difference between $\epsilon_{\rm RP}$ and $\epsilon_{\rm part}$ and the very moderate fluctuations for mid-central collisions in the region of $b=5-9$ fm. This behaviour is directly reflected in the centrality dependent elliptic flow results as it is shown in Section \ref{flow_results}. \section {The Hybrid Approach} To simulate the dynamic evolution of relativistic heavy ion reactions combined microscopic+macroscopic approaches have proven to be very successful in the description of various observables \cite{Nonaka:2006yn,Dumitru:1999sf,Bass:2000ib,Teaney:2000cw,Teaney:2001av,Grassi:2005pm,Andrade:2005tx,Hirano:2005xf,Hirano:2007ei,Andrade:2008xh,Andrade:2008fa}. The approach that we are using here has recently been developed and is based on the UrQMD hadronic transport approach including a (3+1)-dimensional one fluid ideal hydrodynamic evolution \cite{Rischke:1995ir,Rischke:1995mt} for the hot and dense stage of the reaction while the early non-equilibrium stage and the final decoupling is treated in the hadronic cascade \cite{Petersen:2008dd,Petersen:2009vx} \footnote{The code is available as UrQMD-3.3 at http://urqmd.org}. For the present investigation two in principal different setups are employed, but since they are constructed from the very same ingredients fair comparisons can be made. The first configuration is the integrated approach where the whole evolution from the incoming nuclei to final state particle distributions is calculated on an event-by-event-basis. The second possibility is to stop the UrQMD calculation at $t_{\rm start}$ and to average over many events at this time. These averaged initial conditions are then used to calculate the hydrodynamic evolution once and then the Cooper-Frye transition and the subsequent hadronic rescattering is averaged over many events again to obtain good statistics for the observables. In this second setup the spectators are not taken into account for the further evolution. To avoid spoiling the results by this difference in the setting we concentrate on observables at midrapidity. In both setups the particles in the UrQMD initial state are mapped to energy, momentum and net baryon density distributions via three-dimensional Gaussian distributions that represent one particle each \cite{Steinheimer:2007iy}. Two different equations of state are used to exemplify the differences due to this external input. One is a hadron gas equation of state (HG) with the same degrees of freedom as in the UrQMD approach \cite{Zschiesche:2002zr}. The other one is a bag model equation of state (BM) including a strong first order phase transition to the quark gluon plasma with a large latent heat \cite{Rischke:1995mt}. To see if fluctuations in the initial state affect the result differently for different expansion dynamics during the hydrodynamic evolution these two extreme cases have been chosen. The transition from the hydrodynamic evolution to the transport approach when the matter is diluted in the late stage is treated as a gradual transition on an approximated iso-eigentime hyper-surface (see \cite{Li:2008qm} for details). The final rescatterings and resonance decays are taken into account in the hadronic cascade. \section{Elliptic Flow Results} \label{flow_results} The elliptic flow, the second coefficient of the Fourier expansion of the azimuthal distribution of the particles $v_2$, quantifies the momentum anisotropy in the transverse plane \cite{Sorge:1998mk,Ollitrault:1992bk,Bleicher:2000sx}. It is a self-quenching effect, since the coordinate space asymmetry is transformed to a momentum space anisotropy until the system becomes isotropic. The elliptic flow is sensitive to the pressure gradients and therefore to the equation of state of the matter \cite{Stoecker:1986ci,Voloshin:2008dg}. It has also been shown that the elliptic flow is very sensitive to the shear viscosity that is present during the evolution. \begin{figure}[b] \resizebox{0.5\textwidth}{!}{ \centering \includegraphics{v2_imp_eos_sps.eps} } \caption{(Color online) Centrality dependence of elliptic flow of charged particles at midrapidity ($|y|<0.5$) for Pb+Pb collisions at $E_{\rm lab}=40A$ GeV (a) and $E_{\rm lab}=160A$ GeV (b). The horizontal lines indicate the results for averaged initial conditions using the hadron gas EoS (blue full line) and the bag model EoS (black dotted line) while the symbols (full circles for HG-EoS and open squares for BM-EoS) depict the results for the event-by-event calculation.} \label{v2_imp} \end{figure} In Fig. \ref{v2_imp} the centrality dependence for the charged particle elliptic flow is shown. The upper plot (a) presents results from the hybrid model for Pb+Pb collisions at $E_{\rm lab} =40A$ GeV while the lower plot (b) is for the highest SPS energy. The horizontal lines depict calculations for averaged initial conditions, while the symbols represent the full event-by-event-setup. Overall, the differences in the integrated elliptic flow between the two different calculations are small. Within the statistical error bars the averaged results are in line with the results including fluctuations for the corresponding centrality range. In accordance with the results for the impact parameter dependence of the initial eccentricity the fluctuations get larger for peripheral events. The highest elliptic flow values are reached for mid-central collisions ($b=5-11$ fm). Therefore, we will refer to the impact parameter range from $b=5-9$ fm for the following considerations. Furthermore, calculations for the two different equations of state are shown. For the bag model EoS ('BM') the elliptic flow is expected to be smaller than for the hadron gas EoS ('HG') since the speed of sound during the expansion is smaller. This expectation is not fulfilled if one looks at the results \cite{Petersen:2009gu}. The final elliptic flow results for different EoS are very similar and agree within the error bars. The softer expansion seems to be compensated by the longer lifetime of the system. \begin{figure}[t] \resizebox{0.5\textwidth}{!}{ \centering \includegraphics{v2_eos_flucaver_exc.eps} } \caption{(Color online) Excitation function of charged particle elliptic flow for mid-central Au+Au/Pb+Pb collisions in comparison to the experimental data (coloured symbols) \cite{Alt:2003ab,v2data}. Results for the hadron gas EoS with averaged/fluctuating initial conditions are depicted as blue full/broken line while the calculations with the bag model EoS with averaged/fluctuating initial conditions are represented as black dashed/dotted line.} \label{v2_exc} \end{figure} The calculation of the integrated elliptic flow at midrapidity in dependence of the beam energy (see Fig. \ref{v2_exc}) confirms the finding that the results do not depend on the equation of state or on the initial condition setup. In the SPS range the results of the hybrid model calculations are in a reasonable agreement with the experimental data while at lower energies the generated elliptic flow is too high. Here, the inclusion of a mean field in the cascade part of the evolution helps to reproduce the data \cite{Petersen:2006vm}. As it has been shown in \cite{Petersen:2009vx} the elliptic flow at SPS energies might be affected by the transition from hydrodynamics to the hadronic transport model and leaves room for finite viscosities during the hydrodynamic expansion. \begin{figure}[t] \resizebox{0.5\textwidth}{!}{ \centering \includegraphics{flow_phases_eos_exc.eps} } \caption{(Color online) Excitation function of charged particle elliptic flow for mid-central Au+Au/Pb+Pb collisions in comparison to the experimental data (coloured symbols)\cite{Alt:2003ab,v2data}. The initial state elliptic flow is shown as the red dashed line. The dotted lines refer to the end of the hydrodynamic evolution while the full lines represents full hybrid calculations for two different EoS.} \label{v2_phases_exc} \end{figure} One might now conclude, that the elliptic flow has to be generated in the very early non-equilibrium stage and in the final state, if the results are not sensitive to the EoS during the hydrodynamic expansion and do not change if the overall setup is changed - event-by-event vs. averaged conditions. In Fig. \ref{v2_phases_exc} the contributions to the final elliptic flow value from the different stages of the evolution are shown. The elliptic flow in the initial state for the hydrodynamic expansion at $t_{\rm start}$ is compatible with zero at all beam energies. At early times, there is not enough transverse expansion to build up elliptic flow. The dotted lines show the elliptic flow directly after the Cooper-Frye transition at the end of the hydrodynamic expansion and at that time the final elliptic flow is already visible. The full lines correspond to the full calculation and indicates that only a small amount of elliptic flow is generated in the final and dilute stage of the collision. So, one is lead to the conclusion that the main part of the elliptic flow is indeed generated during the ideal hydrodynamic expansion. \begin{figure}[t] \resizebox{0.5\textwidth}{!}{ \centering \includegraphics{v2_ptpi_eos_flucaver_sps.eps} } \caption{(Color online) Transverse momentum dependence of elliptic flow of pions at midrapidity in mid-central Pb+Pb collisions at SPS energies ($E_{\rm lab}=40A$ GeV, (a) and $E_{\rm lab}=160A$ GeV, (b)). The full lines depict results for fluctuating initial conditions while the dotted lines refer to averaged initial conditions for two different EoS.} \label{v2_pion_pt} \end{figure} Let us now investigate the transverse momentum dependence of the elliptic flow results. The differential flow results might be more sensitive to the different setups. Fig. \ref{v2_pion_pt} shows the elliptic flow of pions at SPS energies for the two different equations of state. For the hadron gas equation of state the different setups lead to similar results again, while for the bag model the averaged initial conditions lead to reduced elliptic flow, especially at higher $p_t$. The Bag model equation of state seems to be more responding to the initial energy density fluctuations. If there are fluctuations, the transverse slices that freeze-out early and produce high transverse momentum particles have a longer time to expand because of the higher energy density spots whereas for the averaged conditions the whole system behaves more smoothly. \begin{figure}[t] \resizebox{0.5\textwidth}{!}{ \centering \includegraphics{v2_ptp_eos_flucaver_sps.eps} } \caption{(Color online) Transverse momentum dependence of elliptic flow of protons at midrapidity in mid-central Pb+Pb collisions at SPS energies ($E_{\rm lab}=40A$ GeV, (a) and $E_{\rm lab}=160A$ GeV, (b)). The full lines depict results for fluctuating initial conditions while the dotted lines refer to averaged initial conditions for two different EoS.} \label{v2_proton_pt} \end{figure} In Fig. \ref{v2_proton_pt} the transverse momentum dependence of the elliptic flow of protons at SPS energies is shown. The protons are interesting because they reflect the dynamics of the incoming nucleons and the finite net baryon density. For the protons the elliptic flow results are even less sensitive to either the EoS or the initial condition averaging than for the pions. At both energies all calculated $v_2$ curves are compatible with each other within error bars. The comparison to experimental data is postponed until more reliable results become available (for pions the comparison has been published in \cite{Petersen:2009gu}). \section{Summary and Conclusions} The effect of initial state fluctuations on the finally measurable elliptic flow has been studied within a (3+1)d Boltzmann + hydrodynamics transport approach. The elliptic flow has been calculated as a function of the impact parameter, the beam energy and the transverse momentum for two different equations of state and for averaged initial conditions or a full event-by-event setup. These investigations allow for the conclusion that in mid-central ($b=5-9$ fm) heavy ion collisions the final elliptic flow reaches its maximum and the fluctuations due to the initial state fluctuations are at a minimum. It has been confirmed that most of the $v_2$ is build up during the hydrodynamic stage of the evolution. The final integrated and differential elliptic flow for charged particles at SPS energies seems to be mostly sensitive to viscosity and not so much on the equation of state. Therefore, the use of averaged initial profiles does not contribute to the uncertainties for the extraction of transport properties of hot and dense QCD matter based on viscous hydrodynamic calculations, while other ambiguities e.g. the contribution of the bulk viscosity \cite{Denicol:2009am} and different treatments of relaxation times for a multi-component system \cite{Monnai:2010qp} remain under debate. Still the question if this statement is also valid at higher RHIC energies arises, since our approach in its present form can only be applied up to top SPS energies. \section{Acknowledgements} \label{ack} We are grateful to the Center for the Scientific Computing (CSC) at Frankfurt for the computing resources. The authors thank Dirk Rischke for providing the 1-fluid hydrodynamics code. This work was supported by GSI and BMBF. This work was supported by the Hessian LOEWE initiative through the Helmholtz International Center for FAIR (HIC for FAIR). H.P. acknowledges a Feodor Lynen fellowship of the Alexander von Humboldt foundation. This work was supported in part by U.S. department of Energy grant DE-FG02-05ER41367.
\section{Introduction} \label{sec:intro} Let $(X_k)_{k\ge 1}$ be a sequence of random variables whose distribution $f^\star$ lies in one of a nested family of models $(\mathcal{M}_q)_{q\ge 1}$, indexed (and ordered) by the integers. We define the model order as the smallest index $q^\star$ such that the true distribution $f^\star$ lies in the corresponding model class. The model order typically determines the most parsimonious representation of the true distribution of the underlying model (for example, it might determine the parametrization of the model which has the smallest possible dimension). On the other hand, the model order often has a concrete interpretation in terms of the modelling of the underlying phenomenon (for example, the estimation of the number of clusters in a data set, or the number of regimes in an economic time series). Therefore, the problem of estimating the model order from observed data is of significant practical, as well as theoretical, interest. Of course, a satisfactory solution to this problem must provide an estimation method that does not assume prior knowledge on the underlying unknown distribution $f^\star$. In particular, prior bounds on model order and on parameter sets should be avoided. Yet, in this light, even one of the most widely used model selection criteria---the Bayesian Information Criterion (BIC) of Schwarz---is poorly understood. The chief motivation for the use of BIC (as opposed to other model selection criteria, such as Akaike's Information Criterion) is that it is expected to yield a strongly consistent estimator of the model order. However, as is pointed out by Csisz{\'a}r and Shields \cite{CS00}, almost all existing consistency proofs assume a prior upper bound on the order as well as compactness of the underlying parameter sets. This is hardly satisfactory from the theoretical point of view, and provides little confidence in the basic motivation for this method. More delicate questions, such as the minimal penalty that yields a strongly consistent order estimator in the absence of a prior bound on the order, also remain open (the problem of identifying the minimal penalty, which minimizes the probability of underestimating the model order, was also raised in \cite{CS00}). Characterizing strong consistency of penalized likelihood order estimators hinges on a precise understanding of the pathwise fluctuations of the likelihood ratio statistic $$ \sup_{f\in\mathcal{M}_q}\ell_n(f) - \sup_{f\in\mathcal{M}_{q^\star}}\ell_n(f), $$ as $n\to\infty$, \emph{uniformly} in the model order $q>q^\star$ (here $\ell_n(f)$ is the likelihood of $(X_k)_{k\le n}$ under the distribution $f\in\mathcal{M}_q$). When there is a known upper bound on the order $q^\star\le q_{\rm max}<\infty$ and the model classes $\mathcal{M}_q$ are parametrized by a compact subset of Euclidean space, an upper bound on the pathwise fluctuations can be obtained by classical parametric methods: Taylor expansion of the likelihood and an application of a law of iterated logarithm. This approach forms the basis for most consistency proofs for penalized likelihood order estimators in the literature, for example \cite{HQ79,Nis88,Fin90,Ker00,Cha06}. However, such techniques fail in the absence of a prior upper bound: even though each model class $\mathcal{M}_q$ is finite dimensional, the full model $\mathcal{M}=\bigcup_q\mathcal{M}_q$ is infinite dimensional and, as such, the problem in the absence of a prior upper bound is inherently nonparametric.\footnote{ One of the key issues in this setting is to understand the dependence of the fluctuations of the likelihood ratio statistic on the dimension of the model classes $\mathcal{M}_q$. However, one of the main results of this paper shows that for regular parametric models, the fluctuations of the likelihood ratio statistic uniformly in $q>q^\star$ are dimension independent when a prior upper bound is assumed (cf.\ Remark \ref{rem:dimensionfree}), which is certainly not the case in the absence of a prior upper bound. Therefore, we find that the pathwise fluctuations of the likelihood ratio statistic with and without a prior upper bound are qualitatively different. } When the model classes $\mathcal{M}_q$ are noncompact, one must introduce sieves $\mathcal{M}_q^n \subset\mathcal{M}_q^{n+1}\subset\cdots\subset\mathcal{M}_q$ which complicate the problem further (in this case even the parametric theory remains poorly understood, see \cite{Har85,BC93,LS04}). An entirely different approach based on universal coding theory \cite{Fin90,GB03,CMR05,CGG09} yields pathwise upper bounds on the likelihood ratio statistic that do not require prior bounds on the order or compactness of the models. However, these bounds are far from tight and cannot even establish consistency of BIC, let alone smaller penalties (this appears to be a fundamental limitation of this approach due to Rissanen's theorem, see \cite{Ris86,BRY99}). To our knowledge, the only setting in which the pathwise fluctuations of the likelihood ratio statistic has been studied in the absence of a prior bound on the order is that of higher-order Markov chains, where Csisz{\'a}r and Shields \cite{CS00,Csis02} proved consistency of BIC. The proofs in \cite{CS00,Csis02} use delicate estimates specific to Markov chains, and do not yield minimal penalties. However, it was shown in \cite{Han09} that a sharp bound can be obtained in the Markov chain case using techniques from empirical process theory, the main difficulty being the dependence structure of Markov chains. The aim of this paper is to obtain generally applicable upper and lower bounds on the pathwise fluctuations of the likelihood ratio statistic uniformly in the model order $q>q^\star$, in the case of i.i.d.\ observations $(X_k)_{k\ge 1}$, without a prior bound on the model order and in possibly noncompact parameter spaces. We use empirical process methods as in \cite{Han09}, but the difficulties to be surmounted in the present setting are of a different nature. Though the Markov chain models in \cite{CS00,Csis02,Han09} suffer from a lack of independence, geometrically these models are exceedingly simple: the family of $q$th-order Markov chains endowed with the Hellinger distance is simply a Euclidean ball when viewed in the appropriate parametrization. In contrast, in general order estimation problems, one is often faced with model classes that are geometrically very complex. An important case study that will be considered in this paper are finite mixture models (widely used in practice for clustering), which possess a notoriously complicated non-regular geometry. To obtain sharp bounds in such models, we will develop tools that can be used to obtain local and weighted entropy bounds, required for our pathwise fluctuation theorems, in models with non-regular geometric structure. These results are of independent interest: we are not aware of any existing local entropy results for models that possess a nontrivial geometric structure (the difficulty of obtaining local entropy bounds for mixture models is noted, for example, in \cite{GW00,MM11}). Finally, we will apply our results to establish strong consistency of BIC and to identify minimal penalties for order estimation for general model classes, in the absence of prior bounds on the order and the underlying parameter set. The remainder of this paper is organized as follows. Section \ref{sec:pathwise} introduces the general model under consideration, and states our results on the pathwise fluctuations of the likelihood ratio statistic. Section \ref{sec:entropy} states our general results on local and weighted entropies, and considers also the special case of mixture models. Section \ref{sec:consist} derives the consequences for order estimation. Proofs are given in section \ref{sec:proofs}. \section{Pathwise fluctuations of the likelihood ratio statistic} \label{sec:pathwise} \subsection{Basic setting and notation} \label{sec:setting} Let $(E,\mathcal{E},\mu)$ be a measure space. For each $q,n\ge 1$, let $\mathcal{M}_q^n$ be a given family of strictly positive probability densities with respect to $\mu$ (that is, we assume that $\int f d\mu=1$ and that $f>0$ $\mu$-a.e.\ for every $f\in\mathcal{M}_q^n$). Moreover, we assume that $(\mathcal{M}_q^n)_{q,n\ge 1}$ is a nested family of models in the sense that $\mathcal{M}_q^n\subseteq\mathcal{M}_{q+1}^n$ and $\mathcal{M}_q^n\subseteq\mathcal{M}_q^{n+1}$ for all $q,n\ge 1$. Let $\mathcal{M}_q = \bigcup_{n}\mathcal{M}_q^n$, $\mathcal{M}^n = \bigcup_{q}\mathcal{M}_q^n$, $\mathcal{M}=\bigcup_{q,n}\mathcal{M}_q^n$. Consider an i.i.d.\ sequence of $E$-valued random variables $(X_k)_{k\ge 1}$ whose common distribution under the measure $\mathbf{P}^\star$ is $f^\star d\mu$, where $f^\star\in \mathcal{M}_{q^\star}\backslash\mathop{\mathrm{cl}}\mathcal{M}_{q^\star-1}$ for some $q^\star\ge 1$ (here $\mathop{\mathrm{cl}}\mathcal{M}_{q}$ denotes the $L^1(d\mu)$-closure of $\mathcal{M}_q$). The index $q^\star$ is called the \emph{model order}. Let us define the log-likelihood function $$ \ell_n(f) = \sum_{i=1}^n \log f(X_i),\qquad\quad f\in\mathcal{M}. $$ Evidently $\ell_n(f)$ is the log-likelihood of the i.i.d.\ sequence $(X_k)_{k\le n}$ when $X_k\sim f d\mu$. Our aim is to study the pathwise fluctuations of the likelihood ratio statistic $$ \sup_{f\in\mathcal{M}_q^n}\ell_n(f) - \sup_{f\in\mathcal{M}_{q^\star}^n}\ell_n(f) $$ as $n\to\infty$, \emph{uniformly} over the order parameter $q\ge q^\star$. Pathwise upper and lower bounds on the likelihood ratio statistic are the key ingredient in the study of strong consistency of penalized likelihood order estimators (see section \ref{sec:consist}). \begin{example}[Location mixtures] \label{ex:locmix} The guiding example for our theory, the case of location mixtures, will be studied in detail in sections \ref{sec:mixtures} and \ref{sec:mixorder} below. We presently introduce this example in order to clarify our basic setup. Let $E=\mathbb{R}^d$ (with its Borel $\sigma$-field $\mathcal{E}$) and let $\mu$ be the Lebesgue measure on $\mathbb{R}^d$. We fix a strictly positive probability density $f_0$ with respect to $\mu$, and define $f_\theta(x) = f_0(x-\theta)$ for $x,\theta\in\mathbb{R}^d$. Fix a sequence $T(n)\uparrow\infty$ and define $$ \mathcal{M}_q^n = \left\{ \sum_{i=1}^q\pi_if_{\theta_i}: \pi_i\ge 0,~ \sum_{i=1}^q \pi_i = 1,~ \|\theta_i\|\le T(n) \right\}. $$ Then $\mathcal{M}_q$ is the family of all $q$-component mixtures of translates of the density $f_0$, while $\mathcal{M}_q^n$ is the subset of the mixtures $\mathcal{M}_q$ whose translation parameters $(\theta_i)_{i=1,\ldots,q}$ are restricted to a ball of radius $T(n)$. The number of components $q^\star$ of the true mixture $f^\star\in\mathcal{M}$ can be estimated from observations using the order estimator $$ \hat q_n = \mathop{\mathrm{argmax}}_{q\ge 1}\left\{ \sup_{f\in\mathcal{M}_q^n}\ell_n(f) - \mathrm{pen}(n,q)\right\}. $$ Pathwise control of the likelihood ratio statistic allows us to identify what penalties $\mathrm{pen}(n,q)$ and cutoff sequences $T(n)$ yield strong consistency of $\hat q_n$ (cf.\ section \ref{sec:mixorder}). \end{example} \begin{rem} \label{rem:esssup} To avoid measurability problems and other technical complications, we employ throughout this paper the simplifying convention that all uncountable suprema (such as $\sup_{f\in\mathcal{M}_q^n}\ell_n(f)$) are interpreted as essential suprema with respect to the measure $\mathbf{P}^\star$. In the majority of applications the model classes $\mathcal{M}_q^n$ will be separable, in which case the supremum and essential supremum coincide. \end{rem} In the sequel, we will denote by $\|\cdot\|_p$ the $L^p(f^\star d\mu)$-norm, that is, $\|g\|_p^p = \int |g(x)|^p f^\star(x)\mu(dx)$, and we denote by $\langle f,g\rangle=\int f(x)g(x)f^\star(x)\mu(dx)$ the Hilbert space inner product in $L^2(f^\star d\mu)$. Define the Hellinger distance $$ h(f,g)^2 = \int (\sqrt{f}-\sqrt{g})^2d\mu,\qquad f,g\in\mathcal{M}. $$ It is easily seen that $h(f,f^\star) = \|\sqrt{f/f^\star}-1\|_2$. Finally, we will denote by $\mathcal{N}(\mathcal{Q},\delta)$ for any class of functions $\mathcal{Q}$ and $\delta>0$ the minimal number of brackets of $L^2(f^\star d\mu)$-width $\delta$ needed to cover $\mathcal{Q}$: that is, $\mathcal{N}(\mathcal{Q},\delta)$ is the smallest cardinality $N$ of a collection of pairs of functions $\{g_i^L,g_i^U\}_{i=1,\ldots, N}$ such that $\max_{i\le N}\|g_i^U-g_i^L\|_2\le\delta$ and for every $g\in\mathcal{Q}$ we have $g_i^L\le g\le g_i^U$ pointwise for some $i\le N$. \subsection{Upper bound} We aim to obtain a pathwise upper bound on the likelihood ratio statistic that holds \emph{uniformly} in $q>q^\star$. To this end, define for $q,n\ge 1$ and $\varepsilon>0$ the Hellinger ball $$ \mathcal{H}_q^n(\varepsilon)=\{\sqrt{f/f^\star}: f\in\mathcal{M}_q^n,~h(f,f^\star)\le\varepsilon\}. $$ Note that the definition of $\mathcal{H}_q^n(\varepsilon)$ depends on $f^\star$ (which is fixed throughout the paper). The following result shows that the geometry of the Hellinger balls $\mathcal{H}_q^n(\varepsilon)$ controls the pathwise fluctuations of the likelihood ratio statistic. \begin{thm} \label{thm:upperlil} Suppose that for all $n$ sufficiently large, we have $$ \mathcal{N}(\mathcal{H}_q^n(\varepsilon),\delta) \le \left(\frac{K(n)\varepsilon}{\delta}\right)^{\eta(q)} $$ for all $q\ge q^\star$ and $\delta\le\varepsilon$, with $K(n)\ge 1$ and $\eta(q)\ge q$ increasing functions. Then $$ \limsup_{n\to\infty}\frac{1}{\log K(2n)\vee\log\log n} \sup_{q\ge q^\star} \frac{1}{\eta(q)} \left\{ \sup_{f\in\mathcal{M}_q^n}\ell_n(f) - \sup_{f\in\mathcal{M}_{q^\star}^n}\ell_n(f) \right\} \le C $$ $\mathbf{P}^\star$-a.s., where $C>0$ is a universal constant. \end{thm} The proof of Theorem \ref{thm:upperlil} is given in section \ref{sec:proofupperlil} below. The assumption of Theorem \ref{thm:upperlil} on the entropy of the Hellinger balls $\mathcal{H}_q^n(\varepsilon)$ states, roughly speaking, that the class of densities $\mathcal{M}_q^n$ endowed with the Hellinger distance has the same metric structure as a Euclidean ball of dimension $\eta(q)$ and radius of order $K(n)$, at least locally in a neighborhood of the true density $f^\star$. The effective dimension $\eta(q)$ controls the fluctuations of the likelihood ratio statistic as a function of the model order, while the effective radius $K(n)$ controls the fluctuations as a function of time up to a minimal rate of order $\log\log n$. In the following section we will see that the minimal $\log\log n$ rate is indeed optimal. Let us note that the geometric structure required by Theorem \ref{thm:upperlil} is far from obvious in many cases of practical interest. For example, in the case of finite mixtures, the geometry of the parameter sets corresponding to Hellinger balls is notoriously complex and highly non-regular, but we will nonetheless verify the assumption of Theorem \ref{thm:upperlil} (see section \ref{sec:mixtures}). In order to apply Theorem \ref{thm:upperlil} in such cases, we therefore need to develop tools to establish local entropy bounds in models that possess nontrivial geometric structure. Section \ref{sec:entropy} below is devoted to this problem. \subsection{Lower bound} \label{sec:lowerlil} Throughout this section, we specialize to the case that $\mathcal{M}_q^n=\mathcal{M}_q$ does not depend on $n$ (this implies essentially that $\mathcal{M}_q$ is compact). In this setting, Theorem \ref{thm:upperlil} yields an upper bound of order $\log\log n$ on the pathwise fluctuations of the likelihood ratio statistic. The aim of this section is to obtain a matching lower bound of order $\log\log n$, which shows that the minimal rate in Theorem \ref{thm:upperlil} is essentially optimal. For the purposes of a lower bound, the uniformity in $q$ is irrelevant, so that it suffices to restrict attention to some fixed $q>q^\star$. We will in fact obtain a much stronger result in this case, which completely characterizes the precise pathwise asymptotics of the likelihood ratio statistic for fixed $q$ in sufficiently smooth families. The geometric structure required in the present section is somewhat different than that of Theorem \ref{thm:upperlil}. Instead of Hellinger balls, we consider the classes of weighted densities $\mathcal{D}_q = \{d_f:f\in\mathcal{M}_q,~f\ne f^\star\}$ and $\mathcal{D}=\bigcup_q\mathcal{D}_q$, where $$ d_f = \frac{\sqrt{f/f^\star}-1}{h(f,f^\star)},\qquad f\in\mathcal{M},\quad f\ne f^\star. $$ In addition, we define for $\varepsilon>0$ and $q\ge 1$ the local weighted classes $$ \mathcal{D}_q(\varepsilon)=\{d_f:f\in\mathcal{M}_q,~ 0<h(f,f^\star)\le\varepsilon\},\qquad\quad \mathcal{\bar D}_q=\bigcap_{\varepsilon>0} \mathop{\mathrm{cl}}\mathcal{D}_q(\varepsilon), $$ where the closure $\mathop{\mathrm{cl}}\mathcal{D}_q(\varepsilon)$ is in $L^2(f^\star d\mu)$. Evidently $\mathcal{\bar D}_q$ is the set of all possible limit points of $d_f$ as $h(f,f^\star)\to 0$ in $\mathcal{M}_q$. If the neighborhoods of $\mathcal{\bar D}_q$ are sufficiently rich, such limits can be taken along a continuous path in the following sense. \begin{defn} A point $d\in\mathcal{\bar D}_q$ is called \emph{continuously accessible} if there is a path $(f_t)_{t\in\mbox{}]0,1]}\subset\mathcal{M}_q\backslash\{f^\star\}$ such that the map $t\mapsto h(f_t,f^\star)$ is continuous, $h(f_t,f^\star)\to 0$ as $t\to 0$, and $d_{f_t}\to d$ in $L^2(f^\star d\mu)$ as $t\to 0$. The subset of all continuously accessible points in $\mathcal{\bar D}_q$ will be denoted as $\mathcal{\bar D}_q^c$. \end{defn} We can now formulate the main result of this section. \begin{thm} \label{thm:lowerlil} Let $q^\star\le p<q$. Assume that $$ \int_0^1 \sqrt{\log\mathcal{N}(\mathcal{D}_q,u)}\,du<\infty, $$ and that $|d|\le D$ for all $d\in\mathcal{D}_q$ with $D\in L^{2+\alpha}(f^\star d\mu)$ for some $\alpha>0$. Then \begin{multline*} \limsup_{n\to\infty}\frac{1}{\log\log n}\left\{ \sup_{f\in\mathcal{M}_q}\ell_n(f)- \sup_{f\in\mathcal{M}_p}\ell_n(f)\right\} \ge \mbox{}\\ \sup_{g\in L^2_0(f^\star d\mu)}\left\{ \sup_{f\in\mathcal{\bar D}_q^c}(\langle f,g\rangle)_+^2 -\sup_{f\in\mathcal{\bar D}_p}(\langle f,g\rangle)_+^2 \right\}\quad \mathbf{P}^\star\mbox{-a.s.}, \end{multline*} as well as \begin{multline*} \limsup_{n\to\infty}\frac{1}{\log\log n}\left\{ \sup_{f\in\mathcal{M}_q}\ell_n(f)- \sup_{f\in\mathcal{M}_p}\ell_n(f)\right\} \le \mbox{}\\ \sup_{g\in L^2_0(f^\star d\mu)}\left\{ \sup_{f\in\mathcal{\bar D}_q}(\langle f,g\rangle)_+^2 -\sup_{f\in\mathcal{\bar D}_p^c}(\langle f,g\rangle)_+^2 \right\}\quad \mathbf{P}^\star\mbox{-a.s.}, \end{multline*} where $L^2_0(f^\star d\mu) = \{g\in L^2(f^\star d\mu):\|g\|_2\le 1,~ \langle 1,g\rangle=0\}$. \end{thm} Only the first (lower bound) part of the theorem is needed to conclude optimality of the minimal $\log\log n$ rate in Theorem \ref{thm:upperlil}. Indeed, we will obtain as a corollary the following lower bound counterpart to Theorem \ref{thm:upperlil}. \begin{cor} \label{cor:lowerlil} Suppose there exists $q>q^\star$ such that the following hold. \begin{enumerate} \item There is an envelope function $D:E\to\mathbb{R}$ such that $|d|\le D$ for all $d\in\mathcal{D}_q$ and $D\in L^{2+\alpha}(f^\star d\mu)$ for some $\alpha>0$. Moreover, $ \int_0^1 \sqrt{\log\mathcal{N}(\mathcal{D}_q,u)}\,du<\infty. $ \item $\mathcal{\bar D}_q^c\backslash\mathcal{\bar D}_{q^\star}$ is nonempty. \end{enumerate} Let $\eta(q)>0$ be an arbitrary positive function. Then $$ \limsup_{n\to\infty}\frac{1}{\log\log n} \sup_{q\ge q^\star}\frac{1}{\eta(q)} \left\{ \sup_{f\in\mathcal{M}_q}\ell_n(f) - \sup_{f\in\mathcal{M}_{q^\star}}\ell_n(f) \right\} \ge C_0 $$ $\mathbf{P}^\star$-a.s., where $C_0>0$ is nonrandom but may depend on $f^\star$ and $\eta$. \end{cor} The proofs of Theorem \ref{thm:lowerlil} and Corollary \ref{cor:lowerlil} are given in section \ref{sec:prooflowerlil} below. The fact that the geometric assumptions in Theorem \ref{thm:lowerlil} and Corollary \ref{cor:lowerlil} are expressed in terms of weighted classes is not surprising, as the sharp asymptotic expression provided by Theorem \ref{thm:lowerlil} for the pathwise fluctuations of the likelihood ratio statistic are expressed in terms of a variational problem on the weighted classes. Nonetheless, we are naturally led to ask whether there is any relation between the geometric assumptions imposed in the upper bound Theorem \ref{thm:upperlil} and the lower bound Theorem \ref{thm:lowerlil}, which appear to be quite different at first sight. In section \ref{sec:entropy}, we will show that the global entropy of the weighted class is closely related to local entropy, so that the geometric assumptions for the upper and lower bounds are not too far apart. Beside the fundamental value of this observation, the relation between global and local entropies will prove to be an essential tool in order to verify these geometric assumptions in models with a complicated geometry, such as finite mixture models. \begin{rem} \label{rem:dimensionfree} When $\mathcal{\bar D}_q$ and $\mathcal{\bar D}_p$ each contain an $L^2(f^\star d\mu)$-dense subset of continuously accessible points (which is typically the case in sufficiently smooth models), then Theorem \ref{thm:lowerlil} provides the exact characterization \begin{multline*} \limsup_{n\to\infty}\frac{1}{\log\log n}\left\{ \sup_{f\in\mathcal{M}_q}\ell_n(f)- \sup_{f\in\mathcal{M}_p}\ell_n(f)\right\} = \mbox{}\\ \sup_{g\in L^2_0(f^\star d\mu)}\left\{ \sup_{f\in\mathcal{\bar D}_q}(\langle f,g\rangle)_+^2 -\sup_{f\in\mathcal{\bar D}_p}(\langle f,g\rangle)_+^2 \right\}\quad \mathbf{P}^\star\mbox{-a.s.} \end{multline*} Beside its intrinsic interest, this result has a surprising consequence. In the case that $\mathcal{M}_q$ and $\mathcal{M}_p$ are regular parametric models with $\dim(\mathcal{M}_q)> \dim(\mathcal{M}_p)$, one can choose $g\in\mathcal{\bar D}_q$ which is orthogonal to $\mathcal{\bar D}_p$. As $\mathcal{\bar D}_q,\mathcal{\bar D}_p\subseteq L^2_0(f^\star d\mu)$ (see the proof of Corollary \ref{cor:lowerlil}), it follows easily that in this case the right-hand side of the previous equation display is precisely equal to $1$. In particular, we obtain the curious conclusion that in regular parametric models, the magnitude of the fluctuations of the likelihood ratio statistic does not depend on the dimensions $\dim(\mathcal{M}_q)$ and $\dim(\mathcal{M}_p)$. In contrast, it is well known that in regular parametric models, the likelihood ratio statistic itself converges weakly to a chi-square distribution with $\dim(\mathcal{M}_q)-\dim(\mathcal{M}_p)$. degrees of freedom, so the tails of the distribution of the likelihood ratio statistic do in fact depend strongly on the dimensions $\dim(\mathcal{M}_q)$ and $\dim(\mathcal{M}_p)$. Of course, the dimension independence of the pathwise fluctuations will also cease to hold if we are interested in a result that is uniform in the order $q$, as in Theorem \ref{thm:upperlil}. \end{rem} \section{Entropy bounds} \label{sec:entropy} In section \ref{sec:pathwise}, we obtained pathwise bounds on the fluctuations of the likelihood ratio statistic in terms of the geometry of the underlying model classes. However, we have required two distinct types of geometric conditions: local entropy bounds for classes of densities, and global entropy bounds for classes of weighted densities. In this section, we will show that the latter implies the former under appropriate conditions, so that a suitable global entropy bound for weighted densities suffices for all the results in section \ref{sec:pathwise}. We will subsequently show how the requisite entropy bounds can be obtained for the case of location mixtures (cf.\ Example \ref{ex:locmix}). The latter is significant both as an important application, and as a nontrivial case study in obtaining local entropy bounds in models with a complicated geometry. \subsection{From global entropy to local entropy} We are going to establish that local entropy estimates for a class of densities $\mathcal{M}$ can be obtained from global entropy estimates on the associated weighted class $\mathcal{D}$. To this end, let us consider for the purposes of this section a general class of positive probability densities $\mathcal{M}$ with respect to some reference measure $\mu$, a fixed $f^\star\in\mathcal{M}$, and define the class of weighted densities $\mathcal{D}=\{d_f:f\in\mathcal{M},~ f\ne f^\star\}$. In addition, we define for $\delta>0$ the Hellinger ball $\mathcal{H}(\delta) = \{\sqrt{f/f^\star}:h(f,f^\star)\le\delta\}$. We obtain the following result, whose proof is given in section \ref{sec:prooflocalglobal}. \begin{thm} \label{thm:localglobal} Suppose that there exist $q,C_0\ge 1$ and $\varepsilon_0>0$ such that $$ \mathcal{N}(\mathcal{D},\varepsilon)\le \left(\frac{C_0}{\varepsilon}\right)^q \quad\mbox{for every }\varepsilon\le\varepsilon_0. $$ Let $R \ge \sup_f|d_f|$ be an envelope function such that $\|R\|_2<\infty$. Then $$ \mathcal{N}(\mathcal{H}(\delta),\rho) \le \left( \frac{C_1\delta}{\rho} \right)^{q+1} $$ for all $\delta,\rho>0$ such that $\rho/\delta < 4 \wedge 2\|R\|_2$, where $C_1 = 8C_0(1 \vee \|R\|_2/4\varepsilon_0)$. \end{thm} Of course, in the setting of section \ref{sec:pathwise}, we would apply this result to $\mathcal{M}_q^n$, $\mathcal{D}_q^n$, $\mathcal{H}_q^n(\varepsilon)$ for given $n,q$ instead of to $\mathcal{M}$, $\mathcal{D}$, $\mathcal{H}(\varepsilon)$. \subsection{The entropy of mixtures} \label{sec:mixtures} We now develop the requisite entropy bounds in the case of mixtures (Example \ref{ex:locmix}). In this section, let $\mu$ be the Lebesgue measure on $\mathbb{R}^d$. We fix a strictly positive probability density $f_0$ with respect to $\mu$, and consider mixtures of densities in the class $$ \{ f_\theta:\theta\in\mathbb{R}^d \},\qquad f_\theta(x) = f_0(x-\theta)\quad\forall\,x\in\mathbb{R}^d. $$ In everything that follows we fix a nondegenerate mixture $f^\star$ of the form $$ f^\star=\sum_{i=1}^{q^\star}\pi_i^\star f_{\theta_i^\star}. $$ Nondegenerate means that $\pi_i^\star>0$ for all $i$, and $\theta_i^\star\ne\theta_j^\star$ for all $i\ne j$. Let $\Theta\subset\mathbb{R}^d$ be a bounded parameter set such that $\{\theta^\star_i:i=1,\ldots,q^\star\}\subseteq\Theta$, and denote its diameter by $2T$ (that is, $\Theta$ is included in some closed Euclidean ball of radius $T$). We consider for $q\ge 1$ the family of $q$-mixtures $$ \mathcal{M}_q = \left\{\sum_{i=1}^q\pi_if_{\theta_i}: \pi_i\ge 0,~\sum_{i=1}^q\pi_i=1,~\theta_i\in\Theta\right\}, $$ and define the class of weighted densities as $\mathcal{D}_q=\{d_f:f\in\mathcal{M}_q,~f\ne f^\star\}$. Let \begin{align*} H_{0}(x)&=\sup_{\theta\in\Theta}f_{\theta}(x)/f^{\star}(x),\\ H_{1}(x)&=\sup_{\theta\in\Theta} \max_{i=1,\ldots,d}| \partial f_\theta(x) / \partial\theta^i|/f^{\star}(x),\\ H_{2}(x)&=\sup_{\theta\in\Theta} \max_{i,j=1,\ldots,d}| \partial^{2} f_\theta(x) / \partial\theta^i\partial\theta^j|/f^{\star}(x), \\ H_{3}(x)&= \sup_{\theta\in\Theta}\max_{i,j,k=1,\ldots,d}| \partial^{3} f_\theta(x) / \partial\theta^i\partial\theta^j\partial\theta^k|/f^{\star}(x) \end{align*} when $f_0$ is sufficiently differentiable, and let $\mathcal{M}=\bigcup_{q\ge 1}\mathcal{M}_q$ and $\mathcal{D}=\bigcup_{q\ge 1}\mathcal{D}_q$. \begin{rem} In the setting of Example \ref{ex:locmix}, the parameter set $\Theta= \Theta(n)$ depends on $n$, and we then write $\mathcal{M}_q^n$ instead of $\mathcal{M}_q$, etc. However, as the dependence on $n$ is irrelevant for the entropy computation, we consider a fixed parameter set $\Theta$ in this section, and drop the dependence on $n$ in our notation for simplicity. \end{rem} We can now state the result of this section, whose proof is given in section \ref{sec:proofmainmixtures}. \begin{aspta} The following hold: \begin{enumerate} \item $f_0\in C^3$ and $f_0(x)$, $(\partial f_0/\partial\theta^i)(x)$ vanish as $\|x\|\to\infty$. \item $H_{k}\in L^{4}(f^{\star}d\mu)$ for $k=0,1,2$ and $H_{3}\in L^{2}(f^{\star}d\mu)$. \end{enumerate} \end{aspta} \begin{thm} \label{thm:mainmixtures} Suppose that Assumption A holds. Then there exist constants $C^\star$ and $\delta^\star$, which depend on $d$, $q^\star$ and $f^\star$ but not on $\Theta$, $q$ or $\delta$, such that $$ \mathcal{N}(\mathcal{D}_{q},\delta) \le \left( \frac{ C^\star (T\vee 1)^{1/6} (\|H_0\|_4^4\vee\|H_1\|_4^4\vee\|H_2\|_4^4\vee\|H_3\|_2^2) }{\delta} \right)^{18(d+1)q} $$ for all $q\ge q^\star$, $\delta\le\delta^\star$. Moreover, there is a function $D\in L^4(f^\star d\mu)$ with $$ \|D\|_4 \le K^\star (\|H_0\|_4\vee\|H_1\|_4\vee\|H_2\|_4), $$ where $K^\star$ depends only on $d$ and $f^\star$, such that $|d|\le D$ for all $d\in\mathcal{D}$. \end{thm} Let us note that a key aspect of this result is that the dependence of the entropy bound on the order $q$ and on the parameter set $\Theta$ is essentially explicit (see Example \ref{ex:gaussian} below, for example). However, even for fixed $q$ and $\Theta$, the existence of a polynomial bound on the bracketing number of $\mathcal{D}_q$ is far from obvious (previous claims \cite{Ker00,Cha06,AGM09} that such bracketing numbers are polynomial were stated without proof). Define the Hellinger ball $\mathcal{H}_q(\varepsilon)=\{\sqrt{f/f^\star}:f\in\mathcal{M}_q,~ h(f,f^\star)\le\varepsilon\}$. Using Theorem \ref{thm:localglobal}, we immediately obtain the following result on the local entropy of $\mathcal{M}_q$. \begin{cor} \label{cor:localmixtures} Suppose that Assumption A holds. Then $$ \mathcal{N}(\mathcal{H}_{q}(\varepsilon),\delta) \le \left( \frac{C_\Theta\,\varepsilon}{\delta} \right)^{18(d+1)q+1} $$ for all $q\ge q^\star$ and $\delta/\varepsilon\le 1$, where $$ C_\Theta = L^\star\, (T\vee 1)^{1/6} \, (\|H_0\|_4^4\vee\|H_1\|_4^4\vee\|H_2\|_4^4\vee\|H_3\|_2^2)^{5/4} $$ and $L^\star$ is a constant that depends only on $d$, $q^\star$ and $f^\star$. \end{cor} \begin{example}[Gaussian mixtures] \label{ex:gaussian} Consider mixtures of standard Gaussian densities $f_0(x)=(2\pi)^{-d/2}e^{-\|x\|^2/2}$, and let $\Theta(T) = \{\theta\in\mathbb{R}^d:\|\theta\|\le T\}$. Fix a nondegenerate mixture $f^\star$, and define $T^\star= \max_{i=1,\ldots,q^\star}\|\theta_i^\star\|$. Denote by $\mathcal{H}_q(\varepsilon,T)$ the Hellinger ball associated to the parameter set $\Theta(T)$. Then $$ \mathcal{N}(\mathcal{H}_{q}(\varepsilon,T),\delta) \le \left( \frac{C_1^\star e^{C_2^\star T^2}\varepsilon}{\delta} \right)^{18(d+1)q+1} $$ for all $q\ge q^\star$, $T\ge T^\star$, and $\delta/\varepsilon\le 1$, where $C_1^\star,C_2^\star$ are constants that depend on $d$, $q^\star$ and $f^\star$ only. To prove this, it evidently suffices to show that Assumption A holds and that $\|H_k\|_4$ for $k=0,1,2$ and $\|H_3\|_2$ are of order $e^{CT^2}$. These facts are readily verified by a straightforward but tedious computation. \end{example} \begin{rem} We have not optimized the constants in Theorem \ref{thm:mainmixtures} and Corollary \ref{cor:localmixtures}. In particular, the constant $18$ in the exponent can likely be improved. On the other hand, it is unclear whether the dependence on the diameter of $\Theta$ is optimal. Indeed, if one is only interested in global entropy $\mathcal{N}(\mathcal{H}_q,\delta)$ where $\mathcal{H}_q=\{\sqrt{f/f^\star}:f\in\mathcal{M}_q\}$, then it can be read off from the proof of Theorem \ref{thm:mainmixtures} that the constants in the entropy bound depend on $\|H_0\|_1$ and $\|H_1\|_1$ only, which are easily seen to scale polynomially in $T$ due to the translation invariance of the Lebesgue measure. Therefore, for example in the case of Gaussian mixtures, one can obtain a \emph{global} entropy bound which scales only polynomially as a function of $T$, whereas the above \emph{local} entropy bound scales as $e^{CT^2}$. The behavior of local entropies is much more delicate than that of global entropies, however, and we do not know whether it is possible to obtain a local entropy bound that scales polynomially in $T$. \end{rem} The proof of Theorem \ref{thm:mainmixtures} is long and rather technical. Nonetheless, there are some key ideas underlying the proof, which we aim to briefly explain here. The classical approach to controlling local entropies of a parametric class $\mathcal{G}=\{g_\xi:\xi\in\Xi\}$ with $\Xi\subset\mathbb{R}^d$ is as follows (cf.\ \cite{vdV98}, Example 19.7). Suppose that the square root densities $h_\xi = \sqrt{g_\xi/g_{\xi^\star}}$ satisfy the pointwise Lipschitz condition $$ |h_\xi(x)-h_{\xi'}(x)| \le H(x)\,\tnorm{\xi-\xi'},\qquad \xi,\xi'\in\Xi, $$ where $H$ is a function in $L^2$ and $\tnorm{\cdot}$ is a norm on $\Xi$. Suppose, moreover, that $$ h(g_\xi,g_{\xi^\star}) \ge c\,\tnorm{\xi-\xi^\star},\qquad \xi\in\Xi. $$ Define $\Xi(\varepsilon) = \{\xi\in\Xi:\tnorm{\xi-\xi^\star}\le\varepsilon\}$ and $\mathcal{H}(\varepsilon)=\{h_\xi:\xi\in\Xi,~ h(g_\xi,g_{\xi^\star})\le\varepsilon\}$. If $\tnorm{\xi-\xi'}\le\delta$, then $h_{\xi'}-\delta H\le h_\xi\le h_{\xi'}+\delta H$. Therefore, we can control the local bracketing entropy by $\mathcal{N}(\mathcal{H}(c\varepsilon),2\delta\|H\|_2)\le N(\Xi(\varepsilon),\delta)$, where $N(\Xi(\varepsilon),\delta)$ denotes metric entropy. But the metric entropy of a ball can be controlled by a standard volume comparison argument, yielding $N(\Xi(\varepsilon),\delta)\le ((2\varepsilon+\delta)/\delta)^d$. \begin{figure} \centering \includegraphics[width=\textwidth]{hball} \caption{Let $f_\theta(x)=\sqrt{2/\pi}\,e^{-2(x-\theta)^2}$ and $f^\star = f_{0.5}$, and consider the mixture family $\mathcal{M}_2=\{pf_{\theta_1}+ (1-p)f_{\theta_2}:p,\theta_1,\theta_2\in[0,1]\}$. The plots illustrate (a) the set of parameters $(p,\theta_1,\theta_2)$ corresponding to the Hellinger ball $\{f\in\mathcal{M}_2:h(f,f^\star)\le 0.05\}$; and (b) the set of parameters $\{(p,\theta_1,\theta_2):N(p,\theta_1,\theta_2)\le 0.05\}$ with $N(p,\theta_1,\theta_2) = |p(\theta_1-0.5)+(1-p)(\theta_2-0.5)| +\tfrac{1}{2}p(\theta_1-0.5)^2 +\tfrac{1}{2}(1-p)(\theta_2-0.5)^2$. The two plots are related by the local geometry Theorem \ref{thm:localgeom}, which yields $c^\star N(p,\theta_1,\theta_2)\le h(pf_{\theta_1}+(1-p)f_{\theta_2},f^\star) \le C^\star N(p,\theta_1,\theta_2)$ for all $p,\theta_1,\theta_2\in[0,1]$. } \label{fig:hball} \end{figure} Clearly the above properties require $c\tnorm{\xi-\xi^\star}\le h(g_\xi,g_{\xi^\star})\le \|H\|_2\tnorm{\xi-\xi^\star}$ for all $\xi\in\Xi$. Therefore, such an approach can only work when the class $\mathcal{G}$ endowed with the Hellinger distance has a regular geometry (i.e., equivalent to a subset of a finite dimensional Banach space), at least in a neighborhood of the true parameter. This fails miserably in the case of mixture classes $\mathcal{M}_q$, which possess a highly non-regular geometry in a neighborhood of $f^\star$ when $q>q^\star$. In fact, it is easily seen that $h(f,f^\star)=0$ does not even select a unique set of parameters $(\pi_i,\theta_i)_{i=1,\ldots,q}$, as mixture models are non-identifiable, and consequently the Hellinger balls $\mathcal{H}_q(\varepsilon)$ look nothing like norm-balls when viewed as a subset of the parameters $(\pi_i,\theta_i)_{i=1,\ldots,q}$ (cf.\ Figure \ref{fig:hball}). Thus we are faced with two basic difficulties: \begin{enumerate} \item How does one control the subset of parameters $(\pi_i,\theta_i)_{i=1,\ldots,q}$ corresponding to the Hellinger balls $\mathcal{H}_q(\varepsilon)$? \item How does one control the metric entropy of these sets? \end{enumerate} The resolution of the first problem requires us to develop a precise understanding of the local geometry of mixture classes, which is done in Theorem \ref{thm:localgeom} below. One key consequence of this result, for example, is as follows: one can choose sufficiently small neighborhoods $A_1,\ldots,A_{q^\star}$ of $\theta_1,\ldots,\theta_{q^\star}$, respectively, such that the Hellinger distance $h(f,f^\star)$ is bounded above and below up to a constant by the pseudodistance $$ \sum_{\theta_j\in A_0}\pi_j +\sum_{i=1}^{q^\star} \Bigg\{ \Bigg| \sum_{\theta_j\in A_i}\pi_j -\pi_i^\star\Bigg| \\ \mbox{} + \Bigg\| \sum_{\theta_j\in A_i}\pi_j(\theta_j-\theta_i^\star) \Bigg\| + \frac{1}{2} \sum_{\theta_j\in A_i}\pi_j\|\theta_j-\theta_i^\star\|^2 \Bigg\} $$ (here $f=\sum_{i=1}^q\pi_if_{\theta_i}$ and $A_0=\mathbb{R}^d\backslash (A_1\cup\cdots\cup A_{q^\star})$). This pseudodistance quantifies precisely (and rather intuitively) the set of parameters with density close to $f^\star$, see Figure \ref{fig:hball} for an illustration in the simplest possible case. As for the second problem, we avoid it entirely by exploiting Theorem \ref{thm:localglobal} instead of computing directly the local entropy. Using the local geometry Theorem \ref{thm:localgeom} and Taylor expansion, we can approximate the weighted densities $d_f$ by linear combinations of their first and second derivatives with coefficients in a Euclidean ball. The entropy of the latter is easily estimated by the Lipschitz argument indicated above. However, the details are somewhat intricate: Taylor expansion should only be applied to parameters $\theta_j$ that lie close to some $\theta_i^\star$, which requires careful bookkeeping. The local geometry Theorem \ref{thm:localgeom} and the relation between global entropy of weighted densities and local entropy developed in Theorem \ref{thm:localglobal} are key ideas that allow us to obtain local entropy estimates in a geometrically nontrivial model. Let us note that the restriction to location mixtures is only used in the proof of Theorem \ref{thm:localgeom}. We believe that the same technique is applicable to other classes of mixtures (for example, Poisson mixtures or mixtures of densities in an exponential family) provided that the proof of Theorem \ref{thm:localgeom} can be adapted to this setting. \section{Strongly consistent order estimation} \label{sec:consist} The goal of this section is to apply the results of sections \ref{sec:pathwise} and \ref{sec:entropy} to identify what penalties and cutoffs yield strong consistency of penalized likelihood order estimators. We first develop some general consistency and inconsistency results, and then consider specifically the problem of mixture order estimation. \subsection{Consistency and minimal penalties} In this section we consider the general setting introduced in section \ref{sec:setting}. We now suppose, however, that the true model order $q^\star$ (as well as the true density $f^\star$) is not known, so that we must aim to estimate $q^\star$ from an observation sequence $(X_k)_{k\ge 1}$. To this end, define the \emph{penalized likelihood order estimator} as $$ \hat q_n = \mathop{\mathrm{argmax}}_{q\ge 1}\left\{ \sup_{f\in\mathcal{M}_q^n}\ell_n(f) - \mathrm{pen}(n,q)\right\}, $$ where $\mathrm{pen}(n,q)$ is a penalty function. Our goal is to show that the penalized likelihood order estimator is strongly consistent, that is, $\hat q_n\to q^\star$ as $n\to\infty$ $\mathbf{P}^\star$-a.s., for a suitable choice of the penalty (that does not depend on $q^\star$ or $f^\star$). Let us emphasize that the maximum in the definition of $\hat q_n$ is taken over \emph{all} model orders $q\ge 1$, that is, we do not assume that an a priori upper bound on the order is available, in contrast to most previous work on this topic. We obtain the following general result. \begin{thm} \label{thm:mainconsist} Suppose that for all $n$ sufficiently large, we have $$ \mathcal{N}(\mathcal{H}_q^n(\varepsilon),\delta) \le \left(\frac{K(n)\varepsilon}{\delta}\right)^{\eta(q)} $$ for all $q\ge q^\star$ and $\delta\le\varepsilon$, where $K(n)\ge 1$ and $\eta(q)\ge q$ are increasing functions and we assume that $\log K(n)=o(n)$. Let $\mathrm{pen}(n,q)$ be a penalty such that $$ \lim_{n\to\infty}\sup_{q>q^\star} \frac{\eta(q)\{\log K(2n)\vee \log\log n\}}{\mathrm{pen}(n,q)- \mathrm{pen}(n,q^\star)}=0, \qquad \lim_{n\to\infty}\max_{q<q^\star}\frac{\mathrm{pen}(n,q)}{n}=0, $$ and $\mathrm{pen}(n,q)$ is increasing in $q$. Then $\hat q_n\to q^\star$ as $n\to\infty$ $\mathbf{P}^\star$-a.s. \end{thm} Theorem \ref{thm:mainconsist} is proved in section \ref{sec:proofmainconsist} below. Let us now specialize to the case that $\mathcal{M}_q^n=\mathcal{M}_q$ does not depend on $n$, as in section \ref{sec:lowerlil}. In this case, Theorem \ref{thm:mainconsist} immediately yields the following corollary. \begin{cor} \label{cor:mainconsist} Suppose that for all $q\ge q^\star$ and $\delta\le\varepsilon$ $$ \mathcal{N}(\mathcal{H}_q(\varepsilon),\delta) \le \left(\frac{K\varepsilon}{\delta}\right)^{\eta(q)}, $$ where $K\ge 1$ and $\eta(q)\ge q$ is a strictly increasing function. Define the penalty $$ \mathrm{pen}(n,q) = \eta(q)\,\varpi(n), $$ where $\varpi(n)$ is any function such that $$ \lim_{n\to\infty}\frac{\log\log n}{\varpi(n)}=0, \qquad\quad \lim_{n\to\infty}\frac{\varpi(n)}{n}=0. $$ Then $\hat q_n\to q^\star$ as $n\to\infty$ $\mathbf{P}^\star$-a.s. \end{cor} Corollary \ref{cor:mainconsist} states that, when $\mathcal{M}_q^n= \mathcal{M}_q$ does not depend on $n$, the penalized likelihood order estimator is strongly consistent provided the penalty grows faster than $\log\log n$ and slower than $n$. Clearly the $\log\log n$ rate is the minimal one attainable by applying Theorem \ref{thm:mainconsist}. This raises the question whether the $\log\log n$ rate is indeed minimal, in the sense that smaller penalties yield inconsistent estimators. The following result (which follows easily from Theorem \ref{thm:lowerlil}) shows that this is indeed the case, so that the result of Corollary \ref{cor:mainconsist} is essentially optimal. \begin{cor} \label{cor:inconsist} Suppose there exists $q>q^\star$ such that the following hold. \begin{enumerate} \item There is an envelope function $D:E\to\mathbb{R}$ such that $|d|\le D$ for all $d\in\mathcal{D}_q$ and $D\in L^{2+\alpha}(f^\star d\mu)$ for some $\alpha>0$. Moreover, $ \int_0^1 \sqrt{\log\mathcal{N}(\mathcal{D}_q,u)}\,du<\infty. $ \item $\mathcal{\bar D}_q^c\backslash\mathcal{\bar D}_{q^\star}$ is nonempty. \end{enumerate} Let $\eta(q)>0$ be any strictly increasing function, and define the penalty $$ \mathrm{pen}(n,q) = C\,\eta(q)\,\log\log n. $$ If the constant $C>0$ is chosen sufficiently small, then $\hat q_n\ne q^\star$ infinitely often $\mathbf{P}^\star$-a.s. \end{cor} The proof of Corollary \ref{cor:inconsist} is given in section \ref{sec:proofmainconsist}. Let us note that the proof of Corollary \ref{cor:inconsist} actually shows that $\sup_{f\in\mathcal{M}_q}\ell_n(f)-\mathrm{pen}(n,q)>\sup_{f\in\mathcal{M}_{q^\star}} \ell_n(f)-\mathrm{pen}(n,q^\star)$ infinitely often $\mathbf{P}^\star$-a.s., so the conclusion of Corollary \ref{cor:inconsist} is not altered even if we were to impose a prior upper bound on the order. In conclusion, we have shown that when $\mathcal{M}_q^n=\mathcal{M}_q$ does not depend on $n$, penalties growing faster than $\log\log n$ are consistent while the penalty $C\,\eta(q)\log\log n$ is inconsistent when the constant $C$ is sufficiently small. From the proof of Theorem \ref{thm:mainconsist}, we can also see that the penalty $C\,\eta(q)\log\log n$ is consistent when $C$ is sufficiently large. However, the critical value of $C$ may depend on the unknown parameter $f^\star$, so that this \emph{minimal} penalty may not be implementable. On the other hand, assuming that $\eta(q)$ does not depend on $f^\star$ (as is typically the case), penalties satisfying the assumptions of Theorem \ref{thm:mainconsist} obviously do not depend on the unknown parameter $f^\star$ and therefore define admissible estimators. When $\mathcal{M}_q^n$ depends on $n$, larger penalties may be required to ensure consistency, depending on the growth rate of $K(n)$. \subsection{Mixture order estimation} \label{sec:mixorder} We finally apply the results in the previous section to mixture order estimation. Throughout this section, let $E=\mathbb{R}^d$ and let $\mu$ be the Lebesgue measure on $\mathbb{R}^d$. Fix a strictly positive probability density $f_0$ with respect to $\mu$, and define $$ \mathcal{M}_q^n = \left\{ \sum_{i=1}^q\pi_if_{\theta_i}:\pi_i\ge 0,~\sum_{i=1}^q\pi_i=1,~ \theta_i\in\Theta(n)\right\}, $$ where $f_\theta(x)=f_0(x-\theta)$ and $\cdots\subseteq\Theta(n)\subseteq\Theta(n+1)\subseteq\cdots \subset\mathbb{R}^d$ is an increasing family of bounded subsets of $\mathbb{R}^d$. We fix $f^\star\in\mathcal{M}$ throughout this section. In the following, we consider two separate cases. The first case is that of a compact parameter set, where $\Theta(n)=\Theta$ does not depend on $n$. In this setting, we obtain a general result. Then, we consider the noncompact case in the setting of Gaussian mixtures, and illustrate how Theorem \ref{thm:mainconsist} can be used to obtain consistency results in this case. Let us first consider the case of a compact parameter set. Then we obtain a general consistency result under Assumption A (cf.\ section \ref{sec:mixtures}). \begin{prop} \label{prop:mixconsist} Suppose that the parameter set $\Theta(n)=\Theta$ is a bounded subset of $\mathbb{R}^d$ independent of $n$, and that Assumption A holds. If we choose a penalty of the form $$ \mathrm{pen}(n,q) = q\,\omega(n),\qquad\qquad \lim_{n\to\infty}\frac{\log\log n}{\omega(n)}= \lim_{n\to\infty}\frac{\omega(n)}{n}=0, $$ then $\hat q_n\to q^\star$ as $n\to\infty$ $\mathbf{P}^\star$-a.s. On the other hand, if we choose the penalty $$ \mathrm{pen}(n,q)=C\,q\,\log\log n $$ with a sufficiently small constant $C>0$, then $\hat q_n\ne q^\star$ infinitely often $\mathbf{P}^\star$-a.s. \end{prop} We therefore find that in the setting of location mixtures with a compact parameter set, the minimal penalty is of order $\log\log n$. Moreover, the popular BIC penalty \begin{equation} \label{eq:bic} \mathrm{pen}(n,q) = \frac{dq+q-1}{2}\,\log n \end{equation} yields a strongly consistent mixture order estimator in this setting, without a prior upper bound on the order. The requisite Assumption A is a very mild one, which highlights the broad applicability of this result. However, the assumption of a compact parameter space can be quite restrictive in practice. Let us therefore consider a case where the parameter space is noncompact. For simplicity we restrict our attention to Gaussian mixtures, that is, we choose $f_0(x) = (2\pi)^{-d/2}e^{-\|x\|^2/2}$, and we choose the restricted parameter sets $\Theta(n) = \{\theta\in\mathbb{R}^d:\|\theta\|\le T(n)\}$ for some sequence $T(n)\uparrow\infty$. Our aim is to choose the penalty $\mathrm{pen}(n,q)$ and cutoff $T(n)$ so that the penalized likelihood order estimator is strongly consistent. In this setting, we obtain the following result. \begin{prop} \label{prop:mixconsist2} Let $f_0(x) = (2\pi)^{-d/2}e^{-\|x\|^2/2}$ and $\Theta(n) = \{\theta\in\mathbb{R}^d:\|\theta\|\le T(n)\}$, and choose a penalty of the form $\mathrm{pen}(n,q)=q\,\omega(n)$. If $$ \lim_{n\to\infty}\frac{\log\log n}{\omega(n)}= \lim_{n\to\infty}\frac{\omega(n)}{n}=0,\qquad T(n) = O(\sqrt{\log\log n}), $$ then $\hat q_n\to q^\star$ as $n\to\infty$ $\mathbf{P}^\star$-a.s. On the other hand, the BIC penalty (\ref{eq:bic}) yields a strongly consistent order estimator if $T(n) = o(\sqrt{\log n})$. \end{prop} This result illustrates that our theory can establish consistency of the penalized likelihood mixture order estimator without any prior upper bounds on the model order or the magnitude of the true parameters. Let us note that there is nothing particularly special about the Gaussian case: a similar result can be obtained, in principle, for any mixture distribution, as long as one can obtain suitable estimates on the quantities $\|H_i\|_4$ that appear in Corollary \ref{cor:localmixtures} (see Example \ref{ex:gaussian} for the Gaussian case). The proofs of Propositions \ref{prop:mixconsist} and \ref{prop:mixconsist2} appear in section \ref{sec:proofmixconsist} below. \section{Proofs} \label{sec:proofs} \subsection{Proof of Theorem \ref{thm:upperlil}} \label{sec:proofupperlil} The proof of Theorem \ref{thm:upperlil} is based on the following deviation bound for the log-likelihood ratio. This bound is essentially from \cite{vdG00}, Corollary 7.5, but the additional maximum inside the probability is essential for our purposes. \begin{thm} \label{thm:deviationbound} Let $\mathcal{M}$ be a family of strictly positive probability densities with respect to a reference measure $\mu$, fix some $f^\star\in\mathcal{M}$, and define the Hellinger ball $\mathcal{H}(\varepsilon) = \{\sqrt{f/f^\star}: f\in\mathcal{M},~h(f,f^\star)\le\varepsilon\}$ where $h(f,g)^2 = \int (\sqrt{f}-\sqrt{g})^2d\mu$. Suppose that for some constants $K\ge 1$, $p\ge 1$ and all $\delta\le\varepsilon$ $$ \mathcal{N}(\mathcal{H}(\varepsilon),\delta) \le \left( \frac{K\varepsilon}{\delta} \right)^p, $$ where $\mathcal{N}(\mathcal{H}(\varepsilon),\delta)$ is the minimal number of brackets of $L^2(f^\star d\mu)$-width $\delta$ needed to cover $\mathcal{H}(\varepsilon)$. Let $(X_i)_{i\in\mathbb{N}}$ be i.i.d.\ with distribution $f^\star d\mu$. Then $$ \mathbf{P}\left[ \max_{n\le k\le 2n} \sup_{f\in\mathcal{M}} \sum_{j=1}^k\log\left(\frac{f(X_j)}{f^\star(X_j)}\right) \ge \alpha \right] \le C\,e^{-\alpha/C} $$ for all $\alpha\ge Cp(1+\log K)$ and $n\ge 1$, where $C$ is a universal constant. \end{thm} \begin{proof} Define $\bar f = (f+f^\star)/2$ for any $f\in\mathcal{M}$, and define the empirical process $\nu_n(g) = n^{-1/2}\sum_{k=1}^n\{g(X_k)-\mathbf{E}[g(X_k)]\}$. Using concavity of $\log x$ we have $$ \sum_{j=1}^k\log\left(\frac{f(X_j)}{f^\star(X_j)}\right) \le 2k^{1/2} \nu_k(\log(\bar f/f^\star)) - 2k D(f^\star||\bar f), $$ where $D(f^\star||f)= \int \log(f^\star/f) f^\star d\mu$ is relative entropy. As $D(f^\star||f) \ge h(f,f^\star)^2$ \begin{align*} &\mathbf{P}\left[ \max_{n\le k\le 2n} \sup_{f\in\mathcal{M}} \sum_{j=1}^k\log\left(\frac{f(X_j)}{f^\star(X_j)}\right) \ge \alpha \right] \\ &\le \mathbf{P}\left[ \max_{n\le k\le 2n} \sup_{f\in\mathcal{M}} \{ 2k^{1/2}\nu_k(\log(\bar f/f^\star)) - 2k h(\bar f,f^\star)^2 \} \ge \alpha \right] \\ &\le \sum_{s=0}^S \mathbf{P}\left[ \max_{n\le k\le 2n} \sup_{f\in\mathcal{M}:n h(\bar f,f^\star)^2\le\alpha 2^s} |k^{1/2}\nu_k(\log(\bar f/f^\star))| \ge \alpha 2^{s-1} \right] \\ &\le 3\sum_{s=0}^S \max_{n\le k\le 2n} \mathbf{P}\left[ \sup_{f\in\mathcal{M}:h(\bar f,f^\star)^2\le\alpha 2^s n^{-1}} |\nu_k(\log(\{\bar f/f^\star\}^{1/2}))| \ge \alpha 2^{s-5}n^{-1/2} \right], \end{align*} where $S = \min\{s:\alpha 2^s n^{-1}>2\}$, and we have used Lemma \ref{lem:etemadi} below for the last inequality. The remainder of the proof is identical to that of \cite{vdG00}, Theorem 7.4 provided we show that for $\mathcal{\bar H}(\varepsilon) = \{\sqrt{\bar f/f^\star}: f\in\mathcal{M},~h(\bar f,f^\star)\le\varepsilon\}$ $$ \mathcal{N}(\mathcal{\bar H}(\varepsilon),\delta) \le \left( \frac{2\sqrt{2}K\varepsilon}{\delta} \right)^p. $$ To this end, fix $\delta\le\varepsilon$, and note that $h(f,f^\star)\le 4h(\bar f,f^\star)$ by \cite{vdG00}, Lemma 4.2 so that $\{f\in\mathcal{M}:h(\bar f,f^\star)\le\varepsilon\} \subseteq\{f\in\mathcal{M}:h(f,f^\star)\le 4\varepsilon\}$. By assumption, there exist $N\le (2\sqrt{2}K\varepsilon/\delta)^p$ and functions $g_1,\ldots,g_N, h_1,\ldots,h_N$ such that $\|h_i-g_i\|_2\le\delta\sqrt{2}$ for every $i$, and for every $u\in \mathcal{H}(4\varepsilon)$ there is an $i$ such that $g_i\le u\le h_i$. But for every $f\in\mathcal{M}$ such that $h(\bar f,f^\star)\le\varepsilon$, we then have for some $i$ $$ 2^{-1/2}\sqrt{g_i^2 + 1} \le \sqrt{\bar f/f^\star } \le 2^{-1/2}\sqrt{h_i^2+1}. $$ Moreover, using $|\sqrt{a+c}-\sqrt{b+c}|\le |\sqrt{a}-\sqrt{b}|$ for $a,b,c\ge 0$ we obtain $$ \left\|2^{-1/2}\sqrt{h_i^2+1} - 2^{-1/2}\sqrt{g_i^2 + 1}\right\|_2 \le 2^{-1/2} \|h_i-g_i\|_2 \le \delta. $$ The result now follows directly. \end{proof} The following variant of Etemadi's inequality was used in the proof. The proof follows closely that of the classical Etemadi inequality, see \cite{Bil99}, Appendix M19. \begin{lem} \label{lem:etemadi} Let $\mathcal{Q}$ be a family of measurable functions $f:E\to\mathbb{R}$. Then we have for every $\alpha>0$ and $m,n\in\mathbb{N}$, $m\le n$ $$ \mathbf{P}^\star\left[ \max_{k=m,\ldots,n} \sup_{f\in\mathcal{Q}}|S_k(f)| \ge 3\alpha \right] \le 3\max_{k=m,\ldots,n} \mathbf{P}^\star\left[ \sup_{f\in\mathcal{Q}}|S_k(f)| \ge \alpha \right], $$ where $S_n(f)=n^{1/2}\nu_n(f)$. \end{lem} \begin{proof} Define the stopping time $\tau=\inf\left\{k\ge m:\sup_{f\in\mathcal{Q}}|S_k(f)|\ge 3\alpha \right\}$. Then \begin{multline*} \mathbf{P}^\star\left[ \max_{k=m,\ldots,n} \sup_{f\in\mathcal{Q}}|S_k(f)| \ge 3\alpha \right] = \mathbf{P}^\star[\tau\le n] \\ \mbox{}\le \mathbf{P}^\star\left[ \sup_{f\in\mathcal{Q}}|S_n(f)|\ge\alpha \right] + \sum_{k=m}^{n} \mathbf{P}^\star\left[\tau=k\mbox{ and } \sup_{f\in\mathcal{Q}}|S_n(f)|<\alpha\right]. \end{multline*} But on the event $\{\tau=k\mbox{ and } \sup_{f\in\mathcal{Q}}|S_n(f)|<\alpha\}$, we clearly have $$ 2\alpha \le \sup_{f\in\mathcal{Q}}|S_k(f)| -\sup_{f\in\mathcal{Q}}|S_n(f)| \le \sup_{f\in\mathcal{Q}}|S_k(f)-S_n(f)|. $$ Therefore, we can estimate \begin{align*} &\mathbf{P}^\star\left[ \max_{k=m,\ldots,n} \sup_{f\in\mathcal{Q}}|S_k(f)| \ge 3\alpha \right] \\ &\quad\mbox{}\le \mathbf{P}^\star\left[ \sup_{f\in\mathcal{Q}}|S_n(f)|\ge\alpha \right] + \sum_{k=m}^{n} \mathbf{P}^\star\left[\tau=k\mbox{ and } \sup_{f\in\mathcal{Q}}|S_n(f)-S_k(f)|\ge 2\alpha \right] \\ &\quad\le \mathbf{P}^\star\left[ \sup_{f\in\mathcal{Q}}|S_n(f)|\ge\alpha \right] + \max_{k=m,\ldots,n} \mathbf{P}^\star\left[\sup_{f\in\mathcal{Q}}|S_n(f)-S_k(f)|\ge 2\alpha \right], \end{align*} where we have used that $\sup_{f\in\mathcal{Q}}|S_n(f)-S_k(f)|$ and $\{\tau=k\}$ are independent to obtain the last inequality. The proof is now easily completed. \end{proof} We can now complete the proof of Theorem \ref{thm:upperlil}. \begin{proof}[Proof of Theorem \ref{thm:upperlil}] By assumption, we have $f^\star\in\mathcal{M}_q^n$ for all $q\ge q^\star$ when $n$ is sufficiently large. Then by Theorem \ref{thm:deviationbound}, we have for $n$ sufficiently large $$ \mathbf{P}^\star\left[ \max_{n\le k\le 2n} \sup_{f\in\mathcal{M}_q^{2n}} \{\ell_k(f)-\ell_k(f^\star)\}\ge\alpha \right]\le C\,e^{-\alpha/C} $$ for all $\alpha\ge C\eta(q)(1+\log K(2n))$ and $q\ge q^\star$. Using that $\mathcal{M}_q^k\subseteq\mathcal{M}_q^{2n}$ for $n\le k\le 2n$ and $\ell_k(f^\star)\le\sup_{f\in \mathcal{M}_{q^\star}^k} \ell_k(f)$, we have for $n$ sufficiently large $$ \mathbf{P}^\star\left[ \max_{n\le k\le 2n} \sup_{q\ge q^\star} \frac{1}{\eta(q)} \Bigg\{\sup_{f\in\mathcal{M}_q^k} \ell_k(f)-\sup_{f\in \mathcal{M}_{q^\star}^k}\ell_k(f) \Bigg\}\ge\alpha \right]\le \sum_{q=q^\star}^\infty C\,e^{-\alpha\eta(q)/C} $$ for all $\alpha\ge C(1+\log K(2n))$. Let $\beta(n)$ be an increasing function. Then $$ \mathbf{P}^\star\left[ \max_{2^n\le k\le 2^{n+1}} \frac{1}{\beta(k)} \sup_{q\ge q^\star} \frac{1}{\eta(q)} \Bigg\{\sup_{f\in\mathcal{M}_q^k} \ell_k(f)-\sup_{f\in \mathcal{M}_{q^\star}^k}\ell_k(f) \Bigg\}\ge 2C \right]\le \frac{2C}{n^2} $$ for all $n$ sufficiently large, provided that $\beta(2^n)\ge\log K(2^{n+1}) \vee \log\log 2^n$. The proof is now easily completed using the Borel-Cantelli lemma. \end{proof} \subsection{Proof of Theorem \ref{thm:lowerlil}} \label{sec:prooflowerlil} The proof of Theorem \ref{thm:lowerlil} is based on a sequence of auxiliary results. First, we will need a compact law of iterated logarithm for the Strassen functional $$ I_n(g) = \frac{1}{\sqrt{2n\log\log n}}\sum_{i=1}^n \left\{g(X_i)-\mathbf{E}^\star(g(X_1))\right\}. $$ We state the requisite result for future reference. \begin{thm} \label{thm:clil} Let $\mathcal{Q}$ be a family of measurable functions $f:E\to\mathbb{R}$ with $$ \int_0^1\sqrt{\log\mathcal{N}(\mathcal{Q},u)}\,du<\infty. $$ Then, $\mathbf{P}^\star$-a.s., the sequence $(I_n)_{n\ge 0}$ is relatively compact in $\ell_\infty(\mathcal{Q})$, and its set of cluster points coincides precisely with the set $\mathcal{K} = \{f\mapsto \langle f,g\rangle:g\in L^2_0(f^\star d\mu)\}$. \end{thm} Proofs of this result can be found in \cite{Oss87}, Theorem 4.2 or in \cite{LT89}, Theorem 9. We will also need the following simple result, whose proof is omitted. \begin{lem} \label{lem:maximal} Let $(X_i)_{i\ge 1}$ be an i.i.d.\ sequence of random variables, and suppose $\mathbf{E}[|X_1|^p]<\infty$. Then $n^{-1/p}\max_{i=1,\ldots,n}|X_i|\to 0$ a.s.\ as $n\to\infty$. \end{lem} Finally, we will need the following likelihood inequality that relates the log-likelihood ratio $\ell_n(f)-\ell_n(f^\star)$ to the empirical process. Related inequalities appear in \cite{Gas02,LiuShao03,Cha06}, but the following form is perhaps the most natural. \begin{lem} \label{lem:likineq} For any strictly positive probability density $f\ne f^\star$, we have $$ \ell_n(f)-\ell_n(f^\star) \le |\nu_n(d_f)|^2, $$ where $\nu_n(g) = n^{-1/2}\sum_{k=1}^n\{g(X_k)-\mathbf{E}^\star[g(X_k)]\}$ denotes the empirical process. \end{lem} \begin{proof} Note that $$ h(f,f^\star)^2 = 2 - \int 2\sqrt{ff^\star}\,d\mu = -2\,h(f,f^\star)\,\mathbf{E}^\star(d_f(X_1)). $$ Using $\log(1+x)\le x$, we can estimate \begin{align*} &\ell_n(f)-\ell_n(f^\star) = \sum_{i=1}^n 2\,\log(1+h(f,f^\star)\,d_f(X_i)) \le \sum_{i=1}^n 2\,h(f,f^\star)\,d_f(X_i) \\ &\qquad = 2\,\nu_n(d_f)\,h(f,f^\star)\,\sqrt{n} - h(f,f^\star)^2\,n \le \sup_{p\in\mathbb{R}}\left\{2\,\nu_n(d_f)\,p - p^2 \right\}. \end{align*} The proof is easily completed. \end{proof} We can now obtain the following asymptotic expansion of the log-likelihood, which provides a pathwise counterpart to the weak convergence theory in \cite{Gas02,LiuShao03}. \begin{prop} \label{prop:asympexp} Let $q\ge q^\star$. Assume that $$ \int_0^1 \sqrt{\log\mathcal{N}(\mathcal{D}_q,u)}\,du<\infty, $$ and that $|d|\le D$ for all $d\in\mathcal{D}_q$ with $D\in L^{2+\alpha}(f^\star d\mu)$ for some $\alpha>0$. Then \begin{multline*} \sup_{f\in\mathcal{M}_q(4\sqrt{\log\log n/n})} \left\{ 2\,I_n(d_f)\,h(f,f^\star)\,\sqrt{\frac{2n}{\log\log n}} - h(f,f^\star)^2\,\frac{2n}{\log\log n}\right\} \\ -\frac{1}{\log\log n}\left\{ \sup_{f\in\mathcal{M}_q}\ell_n(f)- \ell_n(f^\star)\right\} \xrightarrow{n\to\infty}0\quad\mathbf{P}^\star\mbox{-a.s.}, \end{multline*} where we have defined $\mathcal{M}_q(\varepsilon)= \{f\in\mathcal{M}_q:h(f,f^\star)\le\varepsilon\}$. \end{prop} \begin{proof} We proceed in several steps. \textbf{Step 1} (\emph{localization}). As $q\ge q^\star$ (hence $f^\star\in\mathcal{M}_q$), clearly $$ \sup_{f\in\mathcal{M}_q}\ell_n(f)-\ell_n(f^\star) = \sup_{f\in\mathcal{M}_q:\ell_n(f)-\ell_n(f^\star)\ge 0} \left\{\ell_n(f)-\ell_n(f^\star)\right\}. $$ Now note that, as in the proof of Lemma \ref{lem:likineq}, $$ \ell_n(f)-\ell_n(f^\star) \le 2\,\nu_n(d_f)\,h(f,f^\star)\,\sqrt{n} - h(f,f^\star)^2\,n. $$ Therefore, we can estimate \begin{align*} & \sup_{f\in\mathcal{M}_q:\ell_n(f)-\ell_n(f^\star)\ge 0} h(f,f^\star) \\ &\qquad\mbox{}\le \sup_{f\in\mathcal{M}_q:\ell_n(f)-\ell_n(f^\star)\ge 0} \left\{ h(f,f^\star) + \frac{\ell_n(f)-\ell_n(f^\star)}{n\,h(f,f^\star)}\right\} \\ &\qquad\mbox{}\le \frac{2}{\sqrt{n}} \sup_{f\in\mathcal{M}_q:\ell_n(f)-\ell_n(f^\star)\ge 0} \nu_n(d_f) \le \sqrt{\frac{8\log\log n}{n}} \sup_{d\in\mathcal{D}_q}I_n(d). \end{align*} Now note that we can estimate $$ \sup_{d\in\mathcal{D}_q}I_n(d) \le \inf_{g\in L^2_0(f^\star d\mu)} \sup_{d\in\mathcal{D}_q}|I_n(d) - \langle d,g\rangle| + \sup_{d\in\mathcal{D}_q}\sup_{g\in L^2_0(f^\star d\mu)} \langle d,g\rangle. $$ The first term on the right converges to zero $\mathbf{P}^\star$-a.s.\ as $n\to\infty$ by Theorem \ref{thm:clil}, while the second term is easily seen to equal $\sup_{d\in\mathcal{D}_q}\|d-\langle 1,d\rangle\|_2\le 1$. Therefore $$ \sup_{f\in\mathcal{M}_q:\ell_n(f)-\ell_n(f^\star)\ge 0} h(f,f^\star) \le (1+\varepsilon)\sqrt{\frac{8\log\log n}{n}} $$ eventually as $n\to\infty$ $\mathbf{P}^\star$-a.s.\ for any $\varepsilon>0$. In particular, we find that $$ \{f\in\mathcal{M}_q:\ell_n(f)-\ell_n(f^\star)\ge 0\} \subseteq \left\{f\in\mathcal{M}_q:h(f,f^\star)\le 4\sqrt{\log\log n/n}\right\} $$ eventually as $n\to\infty$ $\mathbf{P}^\star$-a.s.\ This implies that $\mathbf{P}^\star$-a.s.\ eventually as $n\to\infty$ $$ \sup_{f\in\mathcal{M}_q}\ell_n(f)-\ell_n(f^\star) \le \sup_{f\in\mathcal{M}_q:h(f,f^\star)\le 4\sqrt{\log\log n/n}} \left\{\ell_n(f)-\ell_n(f^\star)\right\}. $$ But the reverse inequality clearly holds for all $n\ge 0$, so that in fact $$ \sup_{f\in\mathcal{M}_q}\ell_n(f)-\ell_n(f^\star) = \sup_{f\in\mathcal{M}_q(4\sqrt{\log\log n/n})} \left\{\ell_n(f)-\ell_n(f^\star)\right\} $$ eventually as $n\to\infty$ $\mathbf{P}^\star$-a.s. \textbf{Step 2} (\emph{Taylor expansion}). Taylor expansion gives $2\log(1+x)=2x-x^2+x^2R(x)$, where $R(x)\to 0$ as $x\to 0$. Thus we can write, for any $f\in\mathcal{M}_q$, \begin{multline*} \ell_n(f)-\ell_n(f^\star) = \sum_{i=1}^n 2\,\log(1+h(f,f^\star)\,d_f(X_i)) = \mbox{} \\ 2\,h(f,f^\star)\sum_{i=1}^n \left\{d_f(X_i) +\frac{1}{2}\,h(f,f^\star)\right\} -h(f,f^\star)^2\sum_{i=1}^n (d_f(X_i))^2 \\ \mbox{} -n\,h(f,f^\star)^2 +h(f,f^\star)^2\sum_{i=1}^n (d_f(X_i))^2 R(h(f,f^\star)\,d_f(X_i)). \end{multline*} Using that $\mathbf{E}^\star(d_f(X_1))=-h(f,f^\star)/2$, we therefore have \begin{multline*} \frac{1}{\log\log n}\left\{ \ell_n(f)-\ell_n(f^\star)\right\} = \mbox{}\\ 2\,I_n(d_f)\,h(f,f^\star)\,\sqrt{\frac{2n}{\log\log n}} - h(f,f^\star)^2\,\frac{2n}{\log\log n}+ R_{f,n}\,\frac{n\,h(f,f^\star)^2}{\log\log n} \end{multline*} where we have defined $$ R_{f,n} = \frac{1}{n}\sum_{i=1}^n \{1-(d_f(X_i))^2\} +\frac{1}{n}\sum_{i=1}^n (d_f(X_i))^2 R(h(f,f^\star)\,d_f(X_i)). $$ It follows easily that \begin{multline*} \Bigg|\sup_{f\in\mathcal{M}_q(4\sqrt{\log\log n/n})} \Bigg\{ 2\,I_n(d_f)\,h(f,f^\star)\,\sqrt{\frac{2n}{\log\log n}} - h(f,f^\star)^2\,\frac{2n}{\log\log n}\Bigg\} \\ -\frac{1}{\log\log n}\left\{ \sup_{f\in\mathcal{M}_q}\ell_n(f)- \ell_n(f^\star)\right\}\Bigg| \\ \mbox{} \le \sup_{f\in\mathcal{M}_q(4\sqrt{\log\log n/n})} |R_{f,n}|\,\frac{n\,h(f,f^\star)^2}{\log\log n} \le 16\sup_{f\in\mathcal{M}_q(4\sqrt{\log\log n/n})}|R_{f,n}| \end{multline*} eventually as $n\to\infty$ $\mathbf{P}^\star$-a.s. \textbf{Step 3} (\emph{end of proof}). We can easily estimate \begin{multline*} \sup_{f\in\mathcal{M}_q(4\sqrt{\log\log n/n})}|R_{f,n}| \le \sup_{f\in\mathcal{M}_q} \left|\frac{1}{n}\sum_{i=1}^n \{(d_f(X_i))^2-1\}\right| \\ \mbox{} + \Bigg( \sup_{|x|\le 4\sqrt{\log\log n/n}\max_{i=1,\ldots,n}D(X_i)} |R(x)|\Bigg)\, \frac{1}{n}\sum_{i=1}^n (D(X_i))^2. \end{multline*} As $\mathcal{N}(\mathcal{D}_q,\delta)<\infty$ for every $\delta>0$, the class $\{d^2:d\in\mathcal{D}_q\}$ can be covered by a finite number of brackets with arbitrary small $L^1(f^\star d\mu)$-norm and is therefore $\mathbf{P}^\star$-Glivenko-Cantelli. Moreover, by construction $\mathbf{E}^\star[(d_f(X_i))^2]=1$ for all $f\in\mathcal{M}_q$. Therefore, the first term in this expression converges to zero as $n\to\infty$ $\mathbf{P}^\star$-a.s. On the other hand, by Lemma \ref{lem:maximal} and the fact that $D\in L^{2+\alpha}(f^\star d\mu)$, we have $\mathbf{P}^\star$-a.s. $$ \sqrt{\log\log n/n}\,\max_{i=1,\ldots,n}D(X_i) = \frac{\sqrt{\log\log n}}{n^{\alpha/2(2+\alpha)}}\, n^{-1/(2+\alpha)}\max_{i=1,\ldots,n}D(X_i) \xrightarrow{n\to\infty}0. $$ Therefore the second term converges to zero also, and the proof is complete. \end{proof} \begin{prop} \label{prop:asympexp1} Let $q\ge q^\star$. Assume that $$ \int_0^1 \sqrt{\log\mathcal{N}(\mathcal{D}_q,u)}\,du<\infty, $$ and that $|d|\le D$ for all $d\in\mathcal{D}_q$ with $D\in L^{2+\alpha}(f^\star d\mu)$ for some $\alpha>0$. Then $$ \liminf_{n\to\infty}\left\{ \sup_{d\in\mathcal{\bar D}_q}(I_n(d))_+^2 -\frac{1}{\log\log n}\left\{ \sup_{f\in\mathcal{M}_q}\ell_n(f)- \ell_n(f^\star)\right\}\right\}\ge 0 \quad\mathbf{P}^\star\mbox{-a.s.} $$ \end{prop} \begin{proof} By Proposition \ref{prop:asympexp}, we have \begin{align*} & \liminf_{n\to\infty}\left\{ \sup_{d\in\mathcal{\bar D}_q}(I_n(d))_+^2 -\frac{1}{\log\log n}\left\{ \sup_{f\in\mathcal{M}_q}\ell_n(f)- \ell_n(f^\star)\right\}\right\} \\ &\qquad \ge \liminf_{n\to\infty}\left\{ \sup_{d\in\mathcal{\bar D}_q}(I_n(d))_+^2 - \sup_{f\in\mathcal{M}_q(4\sqrt{\log\log n/n})} \sup_{p\ge 0} \left\{ 2\,I_n(d_f)\,p - p^2\right\} \right\} \\ &\qquad = \liminf_{n\to\infty}\left\{ \sup_{d\in\mathcal{\bar D}_q}(I_n(d))_+^2 - \sup_{f\in\mathcal{M}_q(4\sqrt{\log\log n/n})} (I_n(d_f))_+^2\right\}. \end{align*} Suppose that the right hand side is negative with positive probability. Then there is an $\varepsilon>0$ and a sequence $\tau_n\uparrow\infty$ of random times such that \begin{equation} \label{eq:coras1} \sup_{d\in\mathcal{\bar D}_q}(I_{\tau_n}(d))_+^2 - \sup_{f\in\mathcal{M}_q(4\sqrt{\log\log \tau_n/\tau_n})} (I_{\tau_n}(d_f))_+^2 \le -\varepsilon\quad \mbox{for all }n \end{equation} with positive probability. We will show that this entails a contradiction. By Theorem \ref{thm:clil} (which can be applied here as $\mathcal{N}(\mathcal{D}_q,\delta)=\mathcal{N}(\mathop{\mathrm{cl}}\mathcal{D}_q,\delta)$ for all $\delta>0$), the process $(I_{\tau_n})_{n\ge 0}$ is $\mathbf{P}^\star$-a.s.\ relatively compact in $\ell_\infty(\mathop{\mathrm{cl}}\mathcal{D}_q)$ with \begin{equation} \label{eq:coras2} \inf_{g\in L^2_0(f^\star d\mu)}\sup_{d\in\mathop{\mathrm{cl}}\mathcal{D}_q} |I_{\tau_n}(d)-\langle d,g\rangle|\xrightarrow{n\to\infty}0 \quad\mathbf{P}^\star\mbox{-a.s.} \end{equation} Then there is a set of positive probability on which (\ref{eq:coras1}) and (\ref{eq:coras2}) hold simultaneously. We now concentrate our attention on a single sample path in this set. For any such path, we can clearly find a further subsequence $\sigma_n\uparrow\infty$ such that $\sup_{d\in\mathop{\mathrm{cl}}\mathcal{D}_q}|I_{\sigma_n}(d)-\langle d,g\rangle|\to 0$ as $n\to\infty$ for some $g\in L^2_0(f^\star d\mu)$. Therefore \begin{multline*} \sup_{d\in\mathop{\mathrm{cl}}\mathcal{D}_q} |(I_{\sigma_n}(d))_+^2-(\langle d,g\rangle)_+^2| \le \sup_{d\in\mathop{\mathrm{cl}}\mathcal{D}_q}|I_{\sigma_n}(d)-\langle d,g\rangle|^2 \\ \mbox{}+ 2\sup_{d\in\mathop{\mathrm{cl}}\mathcal{D}_q}|I_{\sigma_n}(d)-\langle d,g\rangle| \sup_{d\in\mathop{\mathrm{cl}}\mathcal{D}_q}|\langle d,g\rangle| \xrightarrow{n\to\infty}0, \end{multline*} where we have used the elementary estimate $|a_+^2-b_+^2| = |a_+-b_+|(a_++b_+) \le |a_+-b_+|(|a_+-b_+|+2b_+) \le |a-b|(|a-b|+2|b|)$ for any $a,b\in\mathbb{R}$, and the fact that $\sup_{d\in\mathop{\mathrm{cl}}\mathcal{D}_q}|\langle d,g\rangle|\le \sup_{d\in\mathop{\mathrm{cl}}\mathcal{D}_q}\|d\|_2\|g\|_2\le 1$. Thus (\ref{eq:coras1}) gives \begin{multline*} \liminf_{n\to\infty}\left\{ \sup_{d\in\mathcal{\bar D}_q}(\langle d,g\rangle)_+^2 - \sup_{f\in\mathcal{M}_q(4\sqrt{\log\log \sigma_n/\sigma_n})} (\langle d_f,g\rangle)_+^2 \right\} = \mbox{}\\ \liminf_{n\to\infty}\left\{ \sup_{d\in\mathcal{\bar D}_q}(I_{\sigma_n}(d))_+^2 - \sup_{f\in\mathcal{M}_q(4\sqrt{\log\log \sigma_n/\sigma_n})} (I_{\sigma_n}(d_f))_+^2 \right\} \le -\varepsilon. \end{multline*} But as $d\mapsto\langle d,g\rangle$ is continuous in $L^2(f^\star d\mu)$ and $\mathop{\mathrm{cl}}\mathcal{D}_q(4\sqrt{\log\log \sigma_n/\sigma_n})$ is compact in $L^2(f^\star d\mu)$ (which follows from $\mathcal{N}(\mathcal{D}_q,\delta)<\infty$ for all $\delta>0$), we have \begin{multline*} \sup_{f\in\mathcal{M}_q(4\sqrt{\log\log \sigma_n/\sigma_n})} (\langle d_f,g\rangle)_+^2 = \sup_{d\in\mathop{\mathrm{cl}}\mathcal{D}_q(4\sqrt{\log\log \sigma_n/\sigma_n})} (\langle d,g\rangle)_+^2 \xrightarrow{n\to\infty} \mbox{}\\ \sup_{d\in\bigcap_{n\ge 0}\mathop{\mathrm{cl}}\mathcal{D}_q(4\sqrt{\log\log \sigma_n/\sigma_n})}(\langle d,g\rangle)_+^2 = \sup_{d\in\mathcal{\bar D}_q}(\langle d,g\rangle)_+^2. \end{multline*} Thus we have a contradiction, completing the proof. \end{proof} We now obtain a converse to the previous result. \begin{prop} \label{prop:asympexp2} Let $q\ge q^\star$. Assume that $$ \int_0^1 \sqrt{\log\mathcal{N}(\mathcal{D}_q,u)}\,du<\infty, $$ and that $|d|\le D$ for all $d\in\mathcal{D}_q$ with $D\in L^{2+\alpha}(f^\star d\mu)$ for some $\alpha>0$. Then $$ \limsup_{n\to\infty}\Bigg\{ \sup_{d\in\mathcal{\bar D}_q^c}(I_n(d))_+^2 -\frac{1}{\log\log n}\Bigg\{ \sup_{f\in\mathcal{M}_q}\ell_n(f)- \ell_n(f^\star)\Bigg\}\Bigg\}\le 0 \quad\mathbf{P}^\star\mbox{-a.s.} $$ \end{prop} \begin{proof} Suppose that the result does not hold true. By Proposition \ref{prop:asympexp}, there is an $\varepsilon>0$ and a sequence $\tau_n\uparrow\infty$ of random times such that \begin{multline*} \sup_{d\in\mathcal{\bar D}_q^c}(I_{\tau_n}(d))_+^2 - \sup_{f\in\mathcal{M}_q(4\sqrt{\log\log\tau_n/\tau_n})} \Bigg\{ - h(f,f^\star)^2\,\frac{2\tau_n}{\log\log\tau_n} \\ \mbox{} +2\,I_{\tau_n}(d_f)\, h(f,f^\star)\,\sqrt{\frac{2\tau_n}{\log\log\tau_n}}\Bigg\} \ge\varepsilon\quad\mbox{for all }n \end{multline*} with positive probability. Proceeding as in the proof of Proposition \ref{prop:asympexp1}, we can then show that there is a sequence of times $\sigma_n\uparrow\infty$ and some $g\in L^2_0(f^\star d\mu)$ such that \begin{multline*} \limsup_{n\to\infty}\Bigg\{ \sup_{d\in\mathcal{\bar D}_q^c}(\langle d,g\rangle)_+^2 - \sup_{f\in\mathcal{M}_q(4\sqrt{\log\log\sigma_n/\sigma_n})} \Bigg\{ - h(f,f^\star)^2\,\frac{2\sigma_n}{\log\log\sigma_n} \\ \mbox{} +2\,\langle d_f,g\rangle\, h(f,f^\star)\,\sqrt{\frac{2\sigma_n}{\log\log\sigma_n}}\Bigg\} \Bigg\}\ge\varepsilon. \end{multline*} We will show that this entails a contradiction. Let $d_0\in\mathcal{\bar D}_q$ be a continuously accessible point. Then there exists an $\alpha_0>0$ (depending on $d_0$) and a path $(f_\alpha)_{\alpha\in\mbox{}]0,\alpha_0]}$ such that $h(f_\alpha,f^\star)=\alpha$ for all $\alpha\in\mbox{}]0,\alpha_0]$ and $d_{f_\alpha}\to d_0$ in $L^2(f^\star d\mu)$ as $\alpha\to 0$. Now choose the sequence $$ \alpha_n = \{(\langle d_0,g\rangle)_+ + \sigma_n^{-1}\} \sqrt{\frac{\log\log\sigma_n}{2\sigma_n}}. $$ As $(\langle d_0,g\rangle)_+ \le \|d_0\|_2\|g\|_2\le 1$, we clearly have $$ 0<\alpha_n<\alpha_0 \wedge 4\sqrt{\log\log\sigma_n/\sigma_n} $$ for all $n$ sufficiently large. In particular $f_{\alpha_n}\in\mathcal{M}_q(4\sqrt{\log\log\sigma_n/\sigma_n})$, so that \begin{multline*} \sup_{f\in\mathcal{M}_q(4\sqrt{\log\log\sigma_n/\sigma_n})} \Bigg\{ 2\,\langle d_f,g\rangle\, h(f,f^\star)\,\sqrt{\frac{2\sigma_n}{\log\log\sigma_n}} - h(f,f^\star)^2\,\frac{2\sigma_n}{\log\log\sigma_n} \Bigg\} \\ \mbox{}\ge 2\,\langle d_{f_{\alpha_n}},g\rangle\,\{(\langle d_0,g\rangle)_+ + \sigma_{n}^{-1}\} - \{(\langle d_0,g\rangle)_+ + \sigma_n^{-1}\}^2. \end{multline*} Therefore, we have \begin{multline*} \limsup_{n\to\infty}\Bigg\{ \sup_{d\in\mathcal{\bar D}_q^c}(\langle d,g\rangle)_+^2 - \sup_{f\in\mathcal{M}_q(4\sqrt{\log\log\sigma_n/\sigma_n})} \Bigg\{ - h(f,f^\star)^2\,\frac{2\sigma_n}{\log\log\sigma_n} \\ \mbox{} +2\,\langle d_f,g\rangle\, h(f,f^\star)\,\sqrt{\frac{2\sigma_n}{\log\log\sigma_n}}\Bigg\} \Bigg\} \le \sup_{d\in\mathcal{\bar D}_q^c}(\langle d,g\rangle)_+^2 - (\langle d_0,g\rangle)_+^2 \end{multline*} for any continuously accessible element $d_0\in\mathcal{\bar D}_q$. But clearly we can choose $d_0$ to make the right hand side of this expression arbitrarily small. Thus we have the desired contradiction, completing the proof. \end{proof} We can now complete the proof of Theorem \ref{thm:lowerlil}. \begin{proof}[Proof of Theorem \ref{thm:lowerlil}] We obtain separately the lower and upper bounds. \textbf{Lower bound.} By Propositions \ref{prop:asympexp1} and \ref{prop:asympexp2}, we have \begin{multline*} \limsup_{n\to\infty}\frac{1}{\log\log n}\left\{ \sup_{f\in\mathcal{M}_q}\ell_n(f)- \sup_{f\in\mathcal{M}_p}\ell_n(f)\right\} \ge \mbox{}\\ \limsup_{n\to\infty}\left\{ \sup_{d\in\mathcal{\bar D}_q^c}(I_n(d))_+^2- \sup_{d\in\mathcal{\bar D}_p}(I_n(d))_+^2 \right\}\quad\mathbf{P}^\star\mbox{-a.s.} \end{multline*} Now fix any $g\in L^2_0(f^\star d\mu)$. By Theorem \ref{thm:clil} (which applies here as $\mathcal{N}(\mathcal{D}_q,\delta)= \mathcal{N}(\mathop{\mathrm{cl}}\mathcal{D}_q,\delta)\ge \mathcal{N}(\mathcal{\bar D}_q,\delta)$ for all $\delta>0$), there is a sequence $\tau_n\uparrow\infty$ of random times such that $I_{\tau_n}\to\langle\,\cdot\,,g\rangle$ in $\ell_\infty(\mathcal{\bar D}_q)$ $\mathbf{P}^\star$-a.s. Therefore $$ \sup_{d\in\mathcal{\bar D}_q^c}(I_{\tau_n}(d))_+^2- \sup_{d\in\mathcal{\bar D}_p}(I_{\tau_n}(d))_+^2 \xrightarrow{n\to\infty} \sup_{d\in\mathcal{\bar D}_q^c}(\langle d,g\rangle)_+^2- \sup_{d\in\mathcal{\bar D}_p}(\langle d,g\rangle)_+^2 \quad\mathbf{P}^\star\mbox{-a.s.}, $$ so that certainly $$ \limsup_{n\to\infty}\frac{1}{\log\log n}\left\{ \sup_{f\in\mathcal{M}_q}\ell_n(f)- \sup_{f\in\mathcal{M}_p}\ell_n(f)\right\} \ge \sup_{d\in\mathcal{\bar D}_q^c}(\langle d,g\rangle)_+^2- \sup_{d\in\mathcal{\bar D}_p}(\langle d,g\rangle)_+^2 $$ $\mathbf{P}^\star$-a.s. But as this inequality holds for every $g\in L^2_0(f^\star d\mu)$, taking the supremum over $g$ gives the requisite lower bound. \textbf{Upper bound.} By Propositions \ref{prop:asympexp1} and \ref{prop:asympexp2}, we have \begin{multline*} \limsup_{n\to\infty}\frac{1}{\log\log n}\left\{ \sup_{f\in\mathcal{M}_q}\ell_n(f)- \sup_{f\in\mathcal{M}_p}\ell_n(f)\right\} \le \mbox{}\\ \limsup_{n\to\infty}\left\{ \sup_{d\in\mathcal{\bar D}_q}(I_n(d))_+^2- \sup_{d\in\mathcal{\bar D}_p^c}(I_n(d))_+^2 \right\}\quad\mathbf{P}^\star\mbox{-a.s.} \end{multline*} It is elementary that for any $d,d'\in\mathcal{\bar D}_q$ and $g\in L^2_0(f^\star d\mu)$ \begin{align*} &(I_n(d))_+^2-(I_n(d'))_+^2 \\ &\quad\le |(I_n(d))_+^2-(\langle d,g\rangle)_+^2| +|(I_n(d'))_+^2-(\langle d',g\rangle)_+^2| +(\langle d,g\rangle)_+^2-(\langle d',g\rangle)_+^2 \\ &\quad\le 2\sup_{d\in\mathcal{\bar D}_q}|(I_n(d))_+^2-(\langle d,g\rangle)_+^2| +(\langle d,g\rangle)_+^2-(\langle d',g\rangle)_+^2. \end{align*} Taking the supremum over $d\in\mathcal{\bar D}_q$ and the infimum over $d'\in\mathcal{\bar D}_p^c$, we find that \begin{align*} &\sup_{d\in\mathcal{\bar D}_q}(I_n(d))_+^2 -\sup_{d\in\mathcal{\bar D}_p^c}(I_n(d))_+^2 \\ &\quad\le 2\sup_{d\in\mathcal{\bar D}_q}|(I_n(d))_+^2-(\langle d,g\rangle)_+^2| +\sup_{d\in\mathcal{\bar D}_q}(\langle d,g\rangle)_+^2 -\sup_{d\in\mathcal{\bar D}_p^c}(\langle d,g\rangle)_+^2 \\ &\quad\le 2\sup_{d\in\mathcal{\bar D}_q}|(I_n(d))_+^2-(\langle d,g\rangle)_+^2| \\ &\qquad\qquad\qquad\qquad +\sup_{g\in L^2_0(f^\star d\mu)}\left\{ \sup_{d\in\mathcal{\bar D}_q}(\langle d,g\rangle)_+^2 -\sup_{d\in\mathcal{\bar D}_p^c}(\langle d,g\rangle)_+^2\right\}. \end{align*} But as this holds for any $g\in L^2_0(f^\star d\mu)$, we finally obtain \begin{multline*} \sup_{d\in\mathcal{\bar D}_q}(I_n(d))_+^2 -\sup_{d\in\mathcal{\bar D}_p^c}(I_n(d))_+^2 \le 2\inf_{g\in L^2_0(f^\star d\mu)} \sup_{d\in\mathcal{\bar D}_q}|(I_n(d))_+^2-(\langle d,g\rangle)_+^2| \\ \mbox{} +\sup_{g\in L^2_0(f^\star d\mu)}\left\{ \sup_{d\in\mathcal{\bar D}_q}(\langle d,g\rangle)_+^2 -\sup_{d\in\mathcal{\bar D}_p^c}(\langle d,g\rangle)_+^2\right\}. \end{multline*} It follows as in the proof of Proposition \ref{prop:asympexp1} that the first term in this expression converges to zero $\mathbf{P}^\star$-a.s. The requisite upper bound follows immediately. \end{proof} Finally, we now complete the proof of Corollary \ref{cor:lowerlil} \begin{proof}[Proof of Corollary \ref{cor:lowerlil}] It evidently suffices to prove that \begin{equation} \label{eq:inconpf} \Gamma:= \sup_{g\in L^2_0(f^\star d\mu)}\Bigg\{ \sup_{d\in\mathcal{\bar D}_q^c}(\langle d,g\rangle)_+^2 -\sup_{d\in\mathcal{\bar D}_{q^\star}}(\langle d,g\rangle)_+^2 \Bigg\}>0. \end{equation} To this end, note that by direct computation $$ \langle 1,d_f\rangle = \frac{\int \sqrt{ff^\star}\,d\mu-1}{h(f,f^\star)} = -\frac{h(f,f^\star)}{2}. $$ Choose $(f_n)_{n\ge 0}\subset\mathcal{M}_q\backslash\{f^\star\}$ such that $h(f_n,f^\star)\to 0$ and $d_{f_n}\to d_0\in\mathcal{\bar D}_q$, then $$ \langle 1,d_0\rangle = \lim_{n\to\infty}\langle 1,d_{f_n}\rangle = -\lim_{n\to\infty}\frac{h(f_n,f^\star)}{2}=0. $$ Moreover, it is immediate that $\|d_0\|_2\le 1$. We have therefore shown that $\mathcal{\bar D}_q\subset L^2_0(f^\star d\mu)$. Now choose $g\in\mathcal{\bar D}_q^c\backslash\mathcal{\bar D}_{q^\star}$. As $\mathcal{\bar D}_{q^\star}$ is closed, it follows directly that $$ \sup_{d\in\mathcal{\bar D}_q^c}(\langle d,g\rangle)_+^2=1,\qquad \qquad \sup_{d\in\mathcal{\bar D}_{q^\star}}(\langle d,g\rangle)_+^2<1. $$ Therefore (\ref{eq:inconpf}) holds, and the proof is complete. \end{proof} \subsection{Proof of Theorem \ref{thm:localglobal}} \label{sec:prooflocalglobal} \begin{proof}[Proof of Theorem \ref{thm:localglobal}] The assumption implies that $$ \mathcal{N}(\mathcal{D},\varepsilon)\le \left(\frac{C_0}{\varepsilon\wedge\varepsilon_0}\right)^q \quad\mbox{for every }\varepsilon>0. $$ If $\varepsilon<\|R\|_2/4$, then $$ \frac{\varepsilon}{\varepsilon\wedge\varepsilon_0} \le 1\vee \frac{\|R\|_2}{4\varepsilon_0}. $$ Defining $C = C_0(1 \vee \|R\|_2/4\varepsilon_0)$, we find that $$ \mathcal{N}(\mathcal{D},\varepsilon)\le \left(\frac{C}{\varepsilon}\right)^q \quad\mbox{for every }\varepsilon<\|R\|_2/4. $$ The remainder of the proof is devoted to establishing that $$ \mathcal{N}(\mathcal{H}(\delta),\rho) \le \left( \frac{8C\delta}{\rho} \right)^{q+1} $$ for all $\delta,\rho>0$ such that $\rho/\delta < 4\wedge 2\|R\|_2$, which is the desired result. Fix $\varepsilon,\delta>0$ and let $N=\mathcal{N}(\mathcal{D},\varepsilon)$. Then there exist $l_1,u_1,\ldots,l_N,u_N$ such that $\|u_i-l_i\|_2\le\varepsilon$ for all $i$ and for every $f$, there is an $i$ such that $l_i\le d_f\le u_i$. Choose $f$ such that $r^{-n}\delta\le h(f,f^\star)\le r^{-n+1}\delta$ (with $r>1$). Then there is an $i$ so that $$ (r^{-n}l_i\wedge r^{-n+1}l_i)\,\delta + 1 \le \sqrt{f/f^\star} \le (r^{-n}u_i\vee r^{-n+1}u_i)\,\delta + 1. $$ Note that \begin{align*} \|u_i\,r^{-n}\delta-l_i\,r^{-n}\delta\|_2 &\le r^{-n}\delta\varepsilon,\\ \|u_i\,r^{-n+1}\delta-l_i\,r^{-n+1}\delta\|_2 &\le r^{-n+1}\delta\varepsilon,\\ \|u_i\,r^{-n+1}\delta-l_i\,r^{-n}\delta\|_2 &\le (r-1)r^{-n}\delta + r^{-n+1}\delta\varepsilon,\\ \|u_i\,r^{-n}\delta-l_i\,r^{-n+1}\delta\|_2 &\le (r-1)r^{-n}\delta + r^{-n+1}\delta\varepsilon, \end{align*} where the latter two estimates follow from $l_i\le d_f\le u_i$, $\|d_f\|_2=1$, and \begin{align*} (u_i-l_i)\,r^{-n}\delta &\le u_i\,r^{-n+1}\delta-l_i\,r^{-n}\delta - d_f\,(r-1)r^{-n}\delta \le (u_i-l_i)\,r^{-n+1}\delta, \\ (u_i-l_i)\,r^{-n}\delta &\le u_i\,r^{-n}\delta-l_i\,r^{-n+1}\delta + d_f\,(r-1)r^{-n}\delta \le (u_i-l_i)\,r^{-n+1}\delta. \end{align*} As $|a\vee b-c\wedge d|\le |a-c|+|a-d|+|b-c|+|b-d|$, we can estimate $$ \|(r^{-n}u_i\vee r^{-n+1}u_i)\,\delta- (r^{-n}l_i\wedge r^{-n+1}l_i)\,\delta\|_2 \le 2(r-1)r^{-n}\delta + 4r^{-n+1}\delta\varepsilon. $$ Therefore, we have shown that $$ \mathcal{N}(\{\sqrt{f/f^\star}:r^{-n}\delta\le h(f,f^\star)\le r^{-n+1}\delta\}, 2(r-1)r^{-n}\delta + 4r^{-n+1}\delta\varepsilon) \le \mathcal{N}(\mathcal{D},\varepsilon) $$ for arbitrary $\varepsilon,\delta>0$, $r>1$, $n\in\mathbb{N}$. In particular, $$ \mathcal{N}(\{\sqrt{f/f^\star}:r^{-n}\delta\le h(f,f^\star)\le r^{-n+1}\delta\},\rho) \le \mathcal{N}(\mathcal{D},\tfrac{1}{4}r^{n-1}\rho/\delta - \tfrac{1}{2}(1-1/r)) $$ for every $\delta>0$, $r>1$, $n\in\mathbb{N}$, $\rho>2(r-1)r^{-n}\delta$. Note that, by finiteness of the bracketing entropies, we can choose an envelope function $R \ge \sup_f|d_f|$ such that $\|R\|_2<\infty$. Then we evidently have $$ 1-r^{-n}\delta R\le\sqrt{f/f^\star} \le 1+r^{-n}\delta R $$ whenever $h(f,f^\star)\le r^{-n}\delta$. Therefore $$ \mathcal{N}(\{\sqrt{f/f^\star}:h(f,f^\star)\le r^{- \lceil H\rceil}\delta\}, 2r^{-H}\delta\|R\|_2)=1 $$ for all $\delta>0$, $r>1$, $H>0$. Thus we can estimate \begin{align*} &\mathcal{N}(\{\sqrt{f/f^\star}:h(f,f^\star)\le\delta\}, 2r^{-H}\delta\|R\|_2) \\ &\quad\le 1+\sum_{n=1}^{\lceil H\rceil} \mathcal{N}(\{\sqrt{f/f^\star}:r^{-n}\delta\le h(f,f^\star)\le r^{-n+1}\delta\},2r^{-H}\delta\|R\|_2) \\ &\quad\le 1+\sum_{n=1}^{\lceil H\rceil} \mathcal{N}(\mathcal{D},\{r^{n-H-1}\|R\|_2- (1-1/r)\}/2) \end{align*} whenever $\delta>0$, $r>1$, $H>0$ such that $\|R\|_2 > (1-1/r)r^{H}$. In particular, $$ \mathcal{N}(\{\sqrt{f/f^\star}:h(f,f^\star)\le\delta\}, 2r^{-H}\delta\|R\|_2) \le 1+\sum_{n=1}^{\lceil H\rceil} \mathcal{N}(\mathcal{D},r^{n-H-1}\|R\|_2/4) $$ whenever $\delta>0$, $r>1$, $H>0$ such that $\|R\|_2 \ge 2(1-1/r)r^{H}$, where we have used that the bracketing number is a nonincreasing function of the bracket size. Now recall that $$ \mathcal{N}(\mathcal{D},\varepsilon)\le \left(\frac{C}{\varepsilon}\right)^q \quad\mbox{for every }0<\varepsilon<\|R\|_2/4, $$ where $q,C\ge 1$. Thus $$ \mathcal{N}(\{\sqrt{f/f^\star}:h(f,f^\star)\le\delta\}, 2r^{-H}\delta\|R\|_2) \le 1+ \sum_{n=1}^{\lceil H\rceil}r^{-(n-1)q} \left( \frac{8C}{2r^{-H}\|R\|_2} \right)^q $$ whenever $\delta>0$, $r>1$, $H>0$ such that $\|R\|_2 \ge 2(1-1/r)r^{H}$. But $$ \sum_{n=1}^{\lceil H\rceil}r^{-(n-1)q} \le \frac{1}{1-1/r^q} \le \frac{1}{1-1/r} \le \frac{\|R\|_2}{2(1-1/r)r^{H}} \frac{4C}{2r^{-H}\|R\|_2} $$ as $r>1$ and $q,C\ge 1$. We can therefore estimate $$ \mathcal{N}(\{\sqrt{f/f^\star}:h(f,f^\star)\le\delta\}, 2r^{-H}\delta\|R\|_2) \le \frac{\|R\|_2}{2(1-1/r)r^{H}} \left( \frac{8C}{2r^{-H}\|R\|_2} \right)^{q+1} $$ whenever $\delta>0$, $r>1$, $H>0$ such that $\|R\|_2 \ge 2(1-1/r)r^{H}$. We now fix $\delta,\rho>0$ such that $\rho/\delta < 4\wedge 2\|R\|_2$, and choose $$ r = \frac{4}{4-\rho/\delta},\qquad H = \frac{\log(2\|R\|_2\delta/\rho)}{\log r}. $$ Clearly $r>1$ and $H>0$. Moreover, note that our choice of $r$ and $H$ implies that $\|R\|_2=2(1-1/r)r^H$ and $\rho=2r^{-H}\delta\|R\|_2$. We have therefore shown that $$ \mathcal{N}(\{\sqrt{f/f^\star}:h(f,f^\star)\le\delta\}, \rho) \le \left( \frac{8C\delta}{\rho} \right)^{q+1} $$ for all $\delta,\rho>0$ such that $\rho/\delta < 4\wedge 2\|R\|_2$. \end{proof} \subsection{Proof of Theorem \ref{thm:mainmixtures}} \label{sec:proofmainmixtures} \subsubsection{The local geometry of mixtures} Define the Euclidean balls $B(\theta,\varepsilon)=\{\theta'\in\mathbb{R}^d: \|\theta-\theta'\|<\varepsilon\}$, denote by $\langle u,v\rangle$ the inner product of two vectors $u,v\in\mathbb{R}^d$, and denote by $\langle A,u\rangle= \{\langle\theta,u\rangle:\theta\in A\}\subseteq\mathbb{R}$ the inner product of a set $A\subseteq\mathbb{R}^d$ with a vector $u\in\mathbb{R}^d$. \begin{lem} \label{lem:bubble} It is possible to choose a bounded convex neighborhood $A_i$ of $\theta_i^\star$ for every $i=1,\ldots,q^\star$ such that, for some linearly independent family $u_1,\ldots,u_d\in\mathbb{R}^d$, the sets $\{\langle A_i,u_j\rangle:i=1,\ldots,q^\star\}$ are disjoint for every $j=1,\ldots,d$. \end{lem} \begin{proof} We first claim that one can choose linearly independent $u_1,\ldots,u_d$ such that $|\{\langle\theta_i^\star,u_j\rangle: i=1,\ldots,q^\star\}|=q^\star$ for every $j=1,\ldots,d$. Indeed, note that the set $\{u\in\mathbb{R}^d:|\{\langle\theta_i^\star,u\rangle: i=1,\ldots,q^\star\}|<q^\star\}$ is a finite union of $(d-1)$-dimensional hyperplanes, which has Lebesgue measure zero. Therefore, if we draw a rotation matrix $T$ at random from the Haar measure on $\mathrm{SO}(d)$, and let $u_i=Te_i$ for all $i=1,\ldots,d$ where $\{e_1,\ldots,e_d\}$ is the standard Euclidean basis in $\mathbb{R}^d$, then the desired property will hold with unit probability. To complete the proof, it suffices to choose $A_i = B(\theta_i^\star,\varepsilon/4)$ with $\varepsilon = \min_k\min_{i\ne j}|\langle\theta_i^\star- \theta_j^\star,u_k\rangle|$. \end{proof} We now fix once and for all a family of neighborhoods $A_1,\ldots,A_{q^\star}$ that satisfy the conditions of Lemma \ref{lem:bubble}. The precise choice of these sets only affects the constants in the proofs below and is therefore irrelevant to our final result; we only presume that $A_1,\ldots,A_{q^\star}$ remain fixed throughout the proofs. Let us also define $A_0=\mathbb{R}^d\backslash(A_1\cup\cdots\cup A_{q^\star})$. Then $\{A_0,\ldots,A_{q^\star}\}$ partitions the parameter set $\mathbb{R}^d$ in such a way that each bounded element $A_i$, $i=1,\ldots,q^\star$ contains precisely one component of the mixture $f^\star$, while the unbounded element $A_0$ contains no components of $f^\star$. This construction is illustrated in Figure \ref{fig:localgeom}. \begin{figure} \centering \input{figlg} \caption{Illustration of the construction of the sets $A_i$ for a mixture with $d=2$ and $q^\star=3$. The sets $A_i$ are chosen in such a way that their projections on some linearly independent vectors $u_1,u_2$ are disjoint. Note that the choice of $u_1,u_2$ is not arbitrary (e.g., consider the projections on the coordinate axes). } \label{fig:localgeom} \end{figure} Let us define for each finite measure $\lambda$ on $\mathbb{R}^d$ the function $$ f_{\lambda}(x) = \int f_\theta(x)\,\lambda(d\theta). $$ We also define the derivatives $D_1f_\theta(x)\in\mathbb{R}^d$ and $D_2f_\theta(x)\in\mathbb{R}^{d\times d}$ as $$ [D_1f_\theta(x)]_{i} = \frac{\partial}{\partial\theta^i}f_\theta(x), \qquad [D_2f_\theta(x)]_{ij} = \frac{\partial^2}{\partial\theta^i \partial\theta^j}f_\theta(x). $$ Denote by $\mathfrak{P}(A)$ the space of probability measures supported on $A\subseteq\mathbb{R}^d$, and denote by $M_+^d$ the family of all $d\times d$ positive semidefinite (symmetric) matrices. \begin{defn} Let us write \begin{multline*} \mathfrak{D} = \{(\eta,\beta,\rho,\tau,\nu): \eta_1,\ldots,\eta_{q^\star}\in\mathbb{R},~ \beta_1,\ldots,\beta_{q^\star}\in\mathbb{R}^d,~ \rho_1,\ldots,\rho_{q^\star}\in M_+^d,\\ \tau_0,\ldots,\tau_{q^\star}\ge 0,~ \nu_0\in\mathfrak{P}(A_0),\ldots, \nu_{q^\star}\in\mathfrak{P}(A_{q^\star})\}. \end{multline*} Then we define for each $(\eta,\beta,\rho,\tau,\nu)\in\mathfrak{D}$ the function $$ \ell(\eta,\beta,\rho,\tau,\nu) = \tau_0\frac{f_{\nu_0}}{f^\star}+ \sum_{i=1}^{q^\star}\left\{ \eta_i\frac{f_{\theta_i^\star}}{f^\star} +\beta_i^*\frac{D_1f_{\theta_i^\star}}{f^\star} +\mathrm{Tr}\left[\rho_i\frac{D_2f_{\theta_i^\star}}{f^\star} \right] +\tau_i\frac{f_{\nu_i}}{f^\star} \right\}, $$ and the nonnegative quantity \begin{multline*} N(\eta,\beta,\rho,\tau,\nu) = \tau_0+ \sum_{i=1}^{q^\star}|\eta_i+\tau_i| + \sum_{i=1}^{q^\star} \left\|\beta_i+\tau_i\int(\theta-\theta_i^\star) \,\nu_i(d\theta)\right\| + \mbox{} \\ \sum_{i=1}^{q^\star} \mathrm{Tr}[\rho_i] + \sum_{i=1}^{q^\star} \frac{\tau_i}{2}\int\|\theta-\theta_i^\star\|^2\nu_i(d\theta). \end{multline*} \end{defn} We now formulate the key result on the local geometry of the mixture class $\mathcal{M}$. \begin{thm} \label{thm:localgeom} Suppose that \begin{enumerate} \item $f_0\in C^2$ and $f_0(x)$, $D_1f_0(x)$ vanish as $\|x\|\to\infty$. \item $\|[D_1f_0]_i/f^\star\|_1<\infty$ and $\|[D_2f_0]_{ij}/f^\star\|_1<\infty$ for all $i,j=1,\ldots,d$. \end{enumerate} Then there exists a constant $c^\star>0$ such that $$ \|\ell(\eta,\beta,\rho,\tau,\nu)\|_1 \ge c^\star\,N(\eta,\beta,\rho,\tau,\nu)\quad \mbox{for all }(\eta,\beta,\rho,\tau,\nu)\in\mathfrak{D}. $$ \emph{[}The constant $c^\star$ may depend on $f^\star$ and $A_1,\ldots,A_{q^\star}$ but not on $\eta,\beta,\rho,\tau,\nu$.\emph{]} \end{thm} Before we turn to the proof, let us introduce a notion that is familiar in quantum physics. If $(\Omega,\Sigma)$ is a measurable space, call the map $\lambda:\Sigma\to\mathbb{R}^{d\times d}$ a \emph{state}\footnote{ Our terminology is in analogy with the usual notion of a state on the $C^*$-algebra $\mathbb{C}^{d\times d}\otimes C_{\mathbb{C}}(\Omega)$, where $\Omega$ is a compact metric space and $C_{\mathbb{C}}(\Omega)$ is the algebra of complex-valued continuous functions on $\Omega$. Such states are precisely represented by the complex-valued counterpart of our definition. } if \begin{enumerate} \item $A\mapsto[\lambda(A)]_{ij}$ is a signed measure for every $i,j=1,\ldots,d$; \item $\lambda(A)$ is a nonnegative symmetric matrix for every $A\in\Sigma$; \item $\mathrm{Tr}[\lambda(\Omega)]=1$. \end{enumerate} It is easily seen that for any unit vector $\xi\in\mathbb{R}^d$, the map $A\mapsto\langle\xi,\lambda(A)\xi\rangle$ is a sub-probability measure. Moreover, if $\xi_1,\ldots,\xi_d\in\mathbb{R}^d$ are linearly independent, there must be at least one $\xi_i$ such that $\langle\xi_i,\lambda(\Omega)\xi_i\rangle>0$. Finally, let $B\subset\mathbb{R}^d$ be a compact set and let $(\lambda_n)_{n\ge 0}$ be a sequence of states on $B$. Then there exists a subsequence along which $\lambda_n$ converges weakly to some state $\lambda$ on $B$ in the sense that $\int \mathrm{Tr}[M(\theta)\lambda_n(d\theta)] \to \int \mathrm{Tr}[M(\theta)\lambda(d\theta)]$ for every continuous function $M:B\to\mathbb{R}^{d\times d}$. To see this, it suffices to note that we may extract a subsequence such that all matrix elements $[\lambda_n]_{ij}$ converge weakly to a signed measure by the compactness of $B$, and it is evident that the limit must again define a state. \begin{proof}[Proof of Theorem \ref{thm:localgeom}] Suppose that the conclusion of the theorem does not hold. Then there must exist a sequence of coefficients $(\eta^{n},\beta^{n},\rho^{n},\tau^{n},\nu^{n})\in\mathfrak{D}$ with $$ \frac{\|\ell(\eta^{n},\beta^{n},\rho^{n},\tau^{n},\nu^{n})\|_1}{ N(\eta^{n},\beta^{n},\rho^{n},\tau^{n},\nu^{n})} \xrightarrow{n\to\infty}0. $$ Let us fix such a sequence throughout the proof. Applying Taylor's theorem to $u\mapsto f_{\theta_i^\star+u(\theta-\theta_i^\star)}$, we can write for $i=1,\ldots,q^\star$ \begin{align*} & \eta_{i}^{n}\frac{f_{\theta_{i}^\star}}{f^\star} + \beta_{i}^{n*} \frac{D_{1}f_{\theta_{i}^\star}}{f^\star} + \mathrm{Tr}\left[\rho_{i}^{n} \frac{D_{2}f_{\theta_{i}^\star}}{f^\star} \right] + \tau_{i}^{n} \frac{f_{\nu_{i}^{n}}}{f^\star} \\ &\mbox{}= \left(\eta_{i}^{n}+\tau_{i}^{n}\right) \frac{f_{\theta_{i}^\star}}{f^\star} + \left(\beta_{i}^{n}+\tau_{i}^{n} \int (\theta-\theta_{i}^\star) \,\nu_{i}^{n}(d\theta)\right)^* \frac{D_{1}f_{\theta_{i}^\star}}{f^\star} + \mathrm{Tr}\left[\rho_{i}^{n} \frac{D_{2}f_{\theta_{i}^\star}}{f^\star} \right] \\ &\quad\mbox{} + \frac{\tau_{i}^{n}}{2}\int \|\theta-\theta_{i}^\star\|^{2}\, \nu_{i}^{n}(d\theta) \int \mathrm{Tr}\bigg[ \bigg\{\int_{0}^{1} \frac{D_{2}f_{\theta_i^\star+u(\theta-\theta_{i}^\star)}}{f^\star}\, 2(1-u)\,du\bigg\} \,\lambda_{i}^{n}(d\theta)\bigg] \end{align*} where $\lambda_{i}^{n}$ is the state on $A_i$ defined by $$ \int \mathrm{Tr}[M(\theta)\,\lambda_{i}^{n}(d\theta)] = \frac{\int \mathrm{Tr}[M(\theta)\,(\theta-\theta_{i}^\star) (\theta-\theta_{i}^\star)^*]\, \nu_{i}^{n}(d\theta)}{ \int\|\theta-\theta_{i}^\star\|^{2}\,\nu_{i}^{n}(d\theta)} $$ (it is clearly no loss of generality to assume that $\nu_i^n$ has no mass at $\theta_i^\star$ for any $i,n$, so that everything is well defined). We now define the coefficients $$ \displaylines{ a_{i}^{n} = \frac{\eta_{i}^{n} + \tau_{i}^{n}}{ N(\eta^{n},\beta^{n},\rho^{n},\tau^{n},\nu^{n})},\qquad b_{i}^{n} = \frac{\beta_{i}^{n} + \tau_{i}^{n}\int(\theta-\theta_{i}^\star)\,\nu_{i}^{n}(d\theta)}{ N(\eta^{n},\beta^{n},\rho^{n},\tau^{n},\nu^{n})}, \cr c_{i}^{n} = \frac{\rho_{i}^{n}}{ N(\eta^{n},\beta^{n},\rho^{n},\tau^{n},\nu^{n})},\qquad d_{i}^{n} = \frac{\frac{\tau_{i}^{n}}{2} \int\|\theta-\theta_{i}^\star\|^{2}\,\nu_{i}^{n}(d\theta)}{ N(\eta^{n},\beta^{n},\rho^{n},\tau^{n},\nu^{n})} } $$ for $i=1,\ldots,q^\star$, and $$ a_0^n = \frac{\tau_{0}^{n}}{ N(\eta^{n},\beta^{n},\rho^{n},\tau^{n},\nu^{n})}. $$ Note that $$ |a_0^n| + \sum_{i=1}^{q^\star}\left\{|a_{i}^{n}|+\|b_{i}^{n}\| +\mathrm{Tr}[c_{i}^{n}]+|d_{i}^{n}|\right\}=1 $$ for all $n$. We may therefore extract a subsequence such that: \begin{enumerate} \item There exist $a_i\in\mathbb{R}$, $b_i\in\mathbb{R}^d$, $c_i\in M_+^d$, and $a_0,d_i\ge 0$ (for $i=1,\ldots,q^\star$) with $|a_0|+\sum_{i=1}^{q^\star}\left\{|a_{i}|+\|b_{i}\|+\mathrm{Tr}[c_{i}] +|d_{i}|\right\}=1$, such that $a_0^n\to a_0$ and $a_i^n\to a_i$, $b_i^n\to b_i$, $c_i^n\to c_i$, $d_i^n\to d_i$ as $n\to\infty$ for all $i=1,\ldots,q^\star$. \item There exists a sub-probability measure $\nu_0$ supported on $A_0$, such that $\nu_0^n$ converges vaguely to $\nu_0$ as $n\to\infty$. \item There exist states $\lambda_i$ supported on $\mathop{\mathrm{cl}}A_i$ for $i=1,\ldots,q^\star$, such that $\lambda_i^n$ converges weakly to $\lambda_i$ as $n\to\infty$ for every $i=1,\ldots,q^\star$. \end{enumerate} It follows that the functions $\ell(\eta^{n},\beta^{n},\rho^{n},\tau^{n},\nu^{n}) / N(\eta^{n},\beta^{n},\rho^{n},\tau^{n},\nu^{n})$ converge pointwise along this subsequence to the function $h/f^\star$ defined by \begin{multline*} h = a_0\,f_{\nu_0} + \sum_{i=1}^{q^\star}\Bigg\{ a_i\,f_{\theta_{i}^\star} + b_i^*\,D_{1}f_{\theta_{i}^\star} + \mathrm{Tr}[c_i\,D_{2}f_{\theta_{i}^\star}] \\ \mbox{} + d_i\int \mathrm{Tr}\bigg[\bigg\{\int_{0}^{1} D_{2}f_{\theta_i^\star+u(\theta-\theta_{i}^\star)}\, 2(1-u)\,du\bigg\} \,\lambda_i(d\theta)\bigg] \Bigg\}. \end{multline*} But as $\|\ell(\eta^{n},\beta^{n},\rho^{n},\tau^{n},\nu^{n})\|_1 / N(\eta^{n},\beta^{n},\rho^{n},\tau^{n},\nu^{n})\to 0$, we have $\|h/f^\star\|_1=0$ by Fatou's lemma. As $f^\star$ is strictly positive, we must have $h\equiv 0$. To proceed, we need the following lemma. \begin{lem} \label{lem:flem} The Fourier transform $F[h](s):=\int e^{\mathrm{i}\langle x,s\rangle}h(x)dx$ is given by \begin{multline*} F[h](s) = F[f_0](s)\, \Bigg[ a_0\int e^{\mathrm{i}\langle\theta,s\rangle}\,\nu_0(d\theta) + \sum_{i=1}^{q^\star}\Bigg\{ a_i\,e^{\mathrm{i}\langle\theta_{i}^\star,s\rangle} + \mathrm{i}\langle b_i,s\rangle\,e^{\mathrm{i}\langle\theta_{i}^\star,s\rangle} \mbox{} \\ - \langle s,c_i s\rangle\,e^{\mathrm{i}\langle\theta_{i}^\star,s\rangle} - d_i\,e^{\mathrm{i}\langle\theta_{i}^\star,s\rangle} \int \phi(\mathrm{i}\langle\theta-\theta_{i}^\star,s\rangle)\, \langle s,\lambda_i(d\theta)s\rangle \Bigg\}\Bigg] \end{multline*} for all $s\in\mathbb{R}^d$. Here we defined the function $\phi(u) = 2(e^u-u-1)/u^2$. \end{lem} \begin{proof} The $a_i,b_i,c_i$ terms are easily computed using integration by parts. It remains to compute the Fourier transform of the function $$ [\Xi_i(x)]_{jk} = \int \bigg\{\int_{0}^{1} [D_{2}f_{\theta_i^\star+u(\theta-\theta_{i}^\star)}(x)]_{jk}\, 2(1-u)\,du\bigg\} \,[\lambda_i(d\theta)]_{kj}. $$ We begin by noting that \begin{multline*} \int \int \int_{0}^{1} |[D_{2}f_{\theta_i^\star+u(\theta-\theta_{i}^\star)}(x)]_{jk}|\, 2(1-u)\,du\,dx\,|[\lambda_i]_{kj}|(d\theta) = \\ \|[\lambda_i]_{kj}\|_{\rm TV}\int |[D_2f_0(x)]_{jk}|\,dx < \infty. \end{multline*} We may therefore apply Fubini's theorem, giving \begin{align*} F[[\Xi_i]_{jk}](s) &= - F[f_0](s)\, s_j s_k\, e^{\mathrm{i}\langle\theta_i^\star,s\rangle} \int\bigg\{\int_{0}^{1} e^{\mathrm{i} u\langle\theta-\theta_{i}^\star,s\rangle} 2(1-u) du\bigg\} [\lambda_i(d\theta)]_{kj} \\ \mbox{} &= - F[f_0](s)\, s_j s_k\,e^{\mathrm{i}\langle\theta_i^\star,s\rangle} \int \phi(\mathrm{i}\langle\theta-\theta_i^\star,s\rangle)\, [\lambda_i(d\theta)]_{kj}, \end{align*} where we have computed the inner integral using integration by parts. \end{proof} Let $u_1,\ldots,u_d\in\mathbb{R}^d$ be a linearly independent family satisfying the condition of Lemma \ref{lem:bubble}. As $F[h](s)=0$ for all $s\in\mathbb{R}^d$, we obtain $$ \Phi^\ell(\mathrm{i} t) := a_0\,\Phi_0^\ell(\mathrm{i} t)+ \sum_{i=1}^{q^\star} e^{\mathrm{i} t\langle\theta_{i}^\star,u_\ell\rangle} \big\{ a_i + \mathrm{i} t \langle b_i,u_\ell\rangle - t^2 \langle u_\ell,c_i u_\ell\rangle - d_i\, t^2\, \Phi_i^\ell(\mathrm{i} t) \big\} = 0 $$ for all $\ell=1,\ldots,d$ and $t\in[-\iota,\iota]\subset\mathbb{R}$ for some $\iota>0$, where we defined $$ \Phi_i^\ell(\mathrm{i} t) = \int \phi(\mathrm{i} t\langle\theta-\theta_{i}^\star,u_\ell\rangle)\, \langle u_\ell,\lambda_i(d\theta)u_\ell\rangle $$ for $i=1,\ldots,q^\star$, and $$ \Phi_0^\ell(\mathrm{i} t) = \int e^{\mathrm{i} t\langle\theta,u_\ell\rangle}\,\nu_0(d\theta). $$ Indeed, it suffices to note that $F[f_0](0)=1$ and that $s\mapsto F[f_0](s)$ is continuous, so that this claim follows from Lemma \ref{lem:flem} and the fact that $F[f_0](s)$ is nonvanishing in a sufficiently small neighborhood of the origin. As all $\lambda_i$ have compact support, it is easily seen that for every $i=1,\ldots,q^\star$, the function $\Phi_i^\ell(z)$ is defined for all $z\in\mathbb{C}$ by a convergent power series. The function $\Psi^\ell(\mathrm{i} t):=\Phi^\ell(\mathrm{i} t)-a_0\,\Phi_0^\ell(\mathrm{i} t)$ is therefore an entire function with $|\Psi^\ell(z)|\le k_1e^{k_2|z|}$ for some $k_1,k_2>0$ and all $z\in\mathbb{C}$. But as $\Phi^\ell(\mathrm{i} t)=0$ for $t\in[-\iota,\iota]$, it follows from \cite{Luk70}, Theorem 7.2.2 that $a_0\,\Phi_0^\ell(\mathrm{i} t)$ is the Fourier transform of a finite measure with compact support. Thus we may assume without loss of generality that the law of $\langle\theta,u_\ell\rangle$ under the sub-probability $\nu_0$ is compactly supported for every $\ell=1,\ldots,d$, so by linear independence $\nu_0$ must be compactly supported. Therefore, the function $\Phi^\ell(z)$ is defined for all $z\in\mathbb{C}$ by a convergent power series. But as $\Phi^\ell(z)$ vanishes for $z\in\mathrm{i}[-\iota,\iota]$, we must have $\Phi^\ell(z)=0$ for all $z\in\mathbb{C}$, and in particular \begin{equation} \label{eq:laplace} \Phi^\ell(t) = a_0\,\Phi_0^\ell(t)+ \sum_{i=1}^{q^\star} e^{t\langle\theta_{i}^\star,u_\ell\rangle} \big\{ a_i + t \langle b_i,u_\ell\rangle + t^2 \langle u_\ell,c_i u_\ell\rangle + d_i\, t^2\, \Phi_i^\ell(t) \big\} = 0 \end{equation} for all $t\in\mathbb{R}$ and $\ell=1,\ldots,d$. In the remainder of the proof, we argue that (\ref{eq:laplace}) can not hold, thus completing the proof by contradiction. At the heart of our proof is an inductive argument. Recall that by construction, the projections $\{\langle A_i,u_\ell\rangle: i=1,\ldots,q^\star\}$ are disjoint open intervals in $\mathbb{R}$ for every $\ell=1,\ldots,d$. We can therefore relabel them in increasing order: that is, define $(\ell1),\ldots,(\ell q^\star)\in\{1,\ldots,q^\star\}$ so that $\langle \theta_{(\ell1)}^\star,u_\ell\rangle < \langle \theta_{(\ell2)}^\star,u_\ell\rangle < \cdots < \langle \theta_{(\ell q^\star)}^\star,u_\ell\rangle$. The following key result provides the inductive step in our proof. \begin{prop} \label{prop:finduction} Fix $\ell\in\{1,\ldots,d\}$, and define $$ \tilde\Phi_0^\ell(t) := a_0\,\Phi_0^\ell(t)+\sum_{i=1}^{q^\star} a_i\,e^{t\langle\theta_{i}^\star,u_\ell\rangle}. $$ Suppose that for some $j\in\{1,\ldots,q^\star\}$ we have $\Phi^{\ell,j}(t)=0$ for all $t\in\mathbb{R}$, where $$ \Phi^{\ell,j}(t):= \tilde\Phi_0^\ell(t)+ \sum_{i=1}^{j} e^{t\langle\theta_{(\ell i)}^\star,u_\ell\rangle} \big\{ t \langle b_{(\ell i)},u_\ell\rangle + t^2 \langle u_\ell,c_{(\ell i)} u_\ell\rangle + d_{(\ell i)}\, t^2\, \Phi_{(\ell i)}^\ell(t) \big\}. $$ Then $d_{(\ell j)}\langle u_\ell,\lambda_{(\ell j)}(\mathbb{R}^d)u_\ell\rangle=0$, $\langle u_\ell,c_{(\ell j)} u_\ell\rangle=0$, and $\langle b_{(\ell j)},u_\ell\rangle=0$. \end{prop} \begin{proof} Let us write for simplicity $\theta_i^\ell = \langle\theta_i^\star,u_\ell\rangle$, and denote by $\lambda_i^\ell$ and $\nu_0^\ell$ the finite measures on $\mathbb{R}$ defined such that $\int f(x)\lambda_i^\ell(dx) = \int f(\langle\theta,u_\ell\rangle) \langle u_\ell,\lambda_i(d\theta)u_\ell\rangle$ and $\int f(x)\nu_0^\ell(dx) = \int f(\langle\theta,u_\ell\rangle) \nu_0(d\theta)$, respectively. For notational convenience, we will assume in the following that $(\ell i)=i$ and $\nu_0^\ell(\{\theta_i^\ell\})=0$ for all $i=1,\ldots,q^\star$. This entails no loss of generality: the former can always be attained by relabeling of the points $\theta_i^\star$, while $\tilde\Phi_0^\ell$ is unchanged if we replace $\nu_0^\ell$ and $a_i$ by $\nu_0^\ell(\,\cdot\,\cap\mathbb{R}\backslash\{\theta_1^\ell,\ldots, \theta_{q^\star}^\ell\})$ and $a_i + a_0\,\nu_0^\ell(\{\theta_i^\ell\})$, respectively. Note that $$ \langle A_i,u_\ell\rangle = \mbox{}]\theta_i^{\ell-}, \theta_i^{\ell+}[\mbox{}, \quad\mbox{where}\quad \theta_i^{\ell-}<\theta_i^\ell< \theta_i^{\ell+}<\theta_{i+1}^{\ell-}\quad\mbox{for all }i $$ by our assumptions ($\langle A_i,u_\ell\rangle$ must be an interval as $A_i$ is convex). \textbf{Step 1}. We claim that the following hold: $$ a_i = 0\mbox{ for all }i\ge j+1\quad\mbox{and}\quad a_0\,\nu_0^\ell([\theta_{j+1}^\ell,\infty[\mbox{})=0. $$ Indeed, suppose this is not the case. Then it is easily seen that $$ \liminf_{t\to\infty} \frac{|\tilde\Phi_0^\ell(t)|}{ e^{t\theta_{j+1}^\ell}}>0, $$ where we have used that $\nu_0^\ell$ has no mass at $\{\theta_1^\ell,\ldots,\theta_{q^\star}^\ell\}$. On the other hand, as $\phi$ is positive and increasing and as $\lambda_i$ is supported on $\mathop{\mathrm{cl}}A_i$, we can estimate $$ 0\le \frac{t^2\,e^{t\theta_i^\ell}\,\Phi_i^\ell(t)}{e^{t\theta_{j+1}^\ell}} \le t^2\,e^{-t(\theta_{j+1}^\ell-\theta_i^\ell)}\, \phi(t\{\theta_j^{\ell+}-\theta_i^\ell\})\, \lambda_i^\ell(\mathbb{R}) \xrightarrow{t\to\infty}0 $$ for $i=1,\ldots,j$. But then we must have $$ 0 = \liminf_{t\to\infty} \frac{|\Phi^{\ell,j}(t)|}{ e^{t\theta_{j+1}^\ell}}>0, $$ which yields the desired contradiction. \textbf{Step 2}. We claim that the following hold: $$ d_j\lambda_j^\ell([\theta_j^\ell,\infty[\mbox{})=0, \quad \langle u_\ell,c_j u_\ell\rangle = 0,\quad\mbox{and}\quad a_0\,\nu_0^\ell([\theta_j^\ell,\infty[\mbox{})=0. $$ Indeed, suppose this is not the case. As $\nu_0^\ell(\{\theta_j^\ell\})=0$, we can choose $\varepsilon>0$ such that $\nu_0^\ell([\theta_j^\ell+\varepsilon,\infty[\mbox{})\ge \nu_0^\ell([\theta_j^\ell,\infty[\mbox{})/2$. As $a_0,d_j\ge 0$, and using that $\phi$ is positive and increasing with $\phi(0)=1$ and that $e^{\varepsilon t}\ge (\varepsilon t)^2/2$ for $t\ge 0$, we can estimate \begin{multline*} a_0\,\Phi_0^\ell(t)+ e^{t\theta_j^\ell}\big\{ t^2\langle u_\ell,c_j u_\ell\rangle+ d_j\,t^2\,\Phi_j^\ell(t)\big\} \ge \mbox{}\\ t^2\,e^{t\theta_j^\ell}\,\bigg\{ \frac{\varepsilon^2}{4}\, a_0\,\nu_0^\ell([\theta_j^\ell,\infty[\mbox{}) + \langle u_\ell,c_j u_\ell\rangle + d_j\,\lambda_j^\ell([\theta_j^\ell,\infty[\mbox{})\bigg\} > 0 \end{multline*} for all $t\ge 0$. On the other hand, it is easily seen that $$ \frac{1}{t^2\,e^{t\theta_j^\ell}} \left[ \sum_{i=1}^{j} e^{t\theta_i^\ell} \big\{ a_i + t \langle b_i,u_\ell\rangle \big\} + \sum_{i=1}^{j-1} e^{t\theta_i^\ell} \big\{ t^2 \langle u_\ell,c_{i} u_\ell\rangle + d_i\, t^2\, \Phi_i^\ell(t) \big\}\right] \xrightarrow{t\to\infty} 0. $$ But this would imply that $$ 0 = \lim_{t\to\infty} \frac{\Phi^{\ell,j}(t)}{a_0\,\Phi_0^\ell(t)+ e^{t\theta_j^\ell}\{ t^2\langle u_\ell,c_j u_\ell\rangle+ d_j\,t^2\,\Phi_j^\ell(t)\}} = 1, $$ which yields the desired contradiction. \textbf{Step 3}. We claim that the following hold: $$ d_j\,\lambda_j^\ell([\theta_j^{\ell-},\theta_j^\ell[\mbox{})=0 \quad\mbox{and}\quad a_0\,\nu_0^\ell([\theta_j^{\ell-},\theta_j^\ell[\mbox{})=0. $$ Indeed, suppose this is not the case. We can compute \begin{multline*} 0 = \frac{d^2}{dt^2}\left( \frac{\Phi^{\ell,j}(t)}{e^{t\theta_j^\ell}}\right) = d_j\int e^{t(\theta-\theta_j^\ell)}\, \lambda_j^\ell(d\theta) + a_0 \int e^{t(\theta-\theta_j^\ell)}\, (\theta-\theta_j^\ell)^2\,\nu_0^\ell(d\theta) \\ \mbox{} + \sum_{i=1}^{j-1} \frac{d^2}{dt^2}\, e^{-t(\theta_j^\ell-\theta_i^\ell)}\big\{ a_i + t\langle b_i,u_\ell\rangle + t^2\langle u_\ell,c_i u_\ell\rangle + d_i\,t^2\,\Phi_i^\ell(t)\big\}, \end{multline*} where the derivative and integral may be exchanged by \cite{Wil91}, Appendix A16. We now note that as $a_0,d_j\ge 0$, we can estimate for $t\ge 0$ \begin{multline*} d_j\int e^{t(\theta-\theta_j^\ell)}\, \lambda_j^\ell(d\theta) + a_0 \int e^{t(\theta-\theta_j^\ell)}\, (\theta-\theta_j^\ell)^2\,\nu_0^\ell(d\theta) \ge \mbox{}\\ e^{t(\theta_j^{\ell-}-\theta_j^\ell)} \Bigg\{ d_j\,\lambda_j^\ell([\theta_j^{\ell-},\theta_j^\ell[\mbox{}) + a_0\int_{[\theta_j^{\ell-},\theta_j^\ell[} (\theta-\theta_j^\ell)^2\,\nu_0^\ell(d\theta) \Bigg\}>0. \end{multline*} On the other hand, as $(e^x-1)/x$ is positive and increasing, we obtain for $t\ge 0$ \begin{align*} & e^{-t(\theta_j^{\ell-}-\theta_j^\ell)} \left|\frac{d^2}{dt^2}\, e^{-t(\theta_j^\ell-\theta_i^\ell)}\,t^2\,\Phi_i^\ell(t)\right| \\ &\quad \mbox{} = e^{-t(\theta_j^{\ell-}-\theta_j^\ell)} \times e^{-t(\theta_j^\ell-\theta_i^\ell)} \times \Bigg| (\theta_j^\ell-\theta_i^\ell)^2 \int t^2\phi(t\{\theta-\theta_i^\ell\})\,\lambda_i^\ell(d\theta) \\ &\qquad\qquad \mbox{} -2(\theta_j^\ell-\theta_i^\ell) \int \frac{e^{t(\theta-\theta_i^\ell)}-1}{\theta-\theta_i^\ell} \,\lambda_i^\ell(d\theta) + \int e^{t(\theta-\theta_i^\ell)}\,\lambda_i^\ell(d\theta) \Bigg| \\ &\quad \mbox{} \le e^{-t(\theta_j^{\ell-}-\theta_i^\ell)} \Bigg\{ (\theta_j^\ell-\theta_i^\ell)^2\,t^2\, \phi(t\{\theta_i^{\ell+}-\theta_i^\ell\}) \\ &\qquad\qquad\qquad\qquad\qquad \mbox{} + 2\,(\theta_j^\ell-\theta_i^\ell)\, \frac{e^{t(\theta_i^{\ell+}-\theta_i^\ell)}-1}{ \theta_i^{\ell+}-\theta_i^\ell} + e^{t(\theta_i^{\ell+}-\theta_i^\ell)} \Bigg\} \,\lambda_i^\ell(\mathbb{R}), \end{align*} which converges to zero as $t\to\infty$ for every $i<j$. It follows that $$ 0 = \lim_{t\to\infty}\frac{\frac{d^2}{dt^2}\left( \Phi^{\ell,j}(t)/e^{t\theta_j^\ell}\right)}{ d_j\int e^{t(\theta-\theta_j^\ell)}\, \lambda_j^\ell(d\theta) + a_0 \int e^{t(\theta-\theta_j^\ell)}\, (\theta-\theta_j^\ell)^2\,\nu_0^\ell(d\theta)} =1, $$ which yields the desired contradiction. \textbf{Step 4}. Recall that $\lambda_j^\ell$ is supported on $[\theta_j^{\ell-},\theta_j^{\ell+}]$ by construction. We have therefore established in the previous steps that the following hold: $$ d_j\langle u_\ell,\lambda_j(\mathbb{R}^d)u_\ell\rangle= \langle u_\ell,c_j u_\ell\rangle= a_0\,\nu_0^\ell([\theta_j^{\ell-},\infty[\mbox{})=0,\qquad a_i=0\mbox{ for }i>j. $$ It is therefore easily seen that $$ 0 = \lim_{t\to\infty}\frac{\Phi^{\ell,j}(t)}{t\,e^{t\theta_j^\ell}} = \langle b_j,u_\ell\rangle. $$ Thus the proof is complete. \end{proof} We can now perform the induction by starting from (\ref{eq:laplace}) and applying Proposition \ref{prop:finduction} repeatedly. This yields $d_j\langle u_\ell,\lambda_j(\mathbb{R}^d)u_\ell\rangle = \langle u_\ell,c_j u_\ell\rangle = \langle b_j,u_\ell\rangle = 0$ for all $j=1,\ldots,q^\star$ and $\ell=1,\ldots,d$. As $u_1,\ldots,u_d$ are linearly independent and $c_j\in M_+^d$, this implies that $b_j=0$, $c_j=0$ and $d_j=0$ for all $j=1,\ldots,q^\star$, so that $$ a_0\int e^{\mathrm{i}\langle\theta,s\rangle}\,\nu_0(d\theta) + \sum_{i=1}^{q^\star}a_i\,e^{\mathrm{i}\langle\theta_i^\star,s\rangle} = 0 $$ for all $s\in\mathbb{R}^d$ (this follows as above by Lemma \ref{lem:flem}, $h\equiv 0$, $F[f_0](s)\ne 0$ for $s$ in a neighborhood of the origin, and using analyticity). But by the uniqueness of Fourier transforms, this implies that the signed measure $a_0\,\nu_0+\sum_{i=1}^{q^\star} a_i\,\delta_{\{\theta_i^\star\}}$ has no mass. As $\nu_0$ is supported on $A_0$, this implies that $a_j=0$ for all $j=1,\ldots,q^\star$. We have therefore shown that $a_i,b_i,c_i,d_i=0$ for all $i=1,\ldots,q^\star$. But recall that $|a_0|+\sum_{i=1}^{q^\star}\{|a_i|+\|b_i\|+\mathrm{Tr}[c_i]+|d_i|\}=1$, so that evidently $a_0=1$. To complete the proof, it remains to note that $$ \int \frac{\ell(\eta^{n},\beta^{n},\rho^{n},\tau^{n},\nu^{n})}{ N(\eta^{n},\beta^{n},\rho^{n},\tau^{n},\nu^{n})}\,f^\star d\mu = \sum_{i=0}^{q^\star} a_i^n \xrightarrow{n\to\infty} 1. $$ But this is impossible, as $$ \left\| \frac{\ell(\eta^{n},\beta^{n},\rho^{n},\tau^{n},\nu^{n})}{ N(\eta^{n},\beta^{n},\rho^{n},\tau^{n},\nu^{n})}\right\|_1 \xrightarrow{n\to\infty}0 $$ by construction. Thus we have the desired contradiction. \end{proof} \subsubsection{Proof of Theorem \ref{thm:mainmixtures}} The proof of Theorem \ref{thm:mainmixtures} consists of a sequence of approximations, which we develop in the form of lemmas. \emph{Throughout this section, we always presume that Assumption A holds.} We begin by establishing the existence of an envelope function. \begin{lem} \label{lem:senvelope} Define $S = (H_0+H_1+H_2)\,d/c^\star$. Then $S\in L^4(f^\star d\mu)$, and $$ \frac{|f/f^\star-1|}{\|f/f^\star-1\|_1} \le S \quad\mbox{for all }f\in\mathcal{M}. $$ \end{lem} \begin{proof} That $S\in L^4(f^\star d\mu)$ follows directly from Assumption A. To proceed, let $f\in\mathcal{M}_q$, so that we can write $f=\sum_{i=1}^q \pi_i f_{\theta_i}$. Then $$ \frac{f - f^\star}{f^\star} = \sum_{j:\theta_j\in A_0}\pi_j \frac{f_{\theta_j}}{f^\star}+ \sum_{i=1}^{q^\star}\Bigg\{ \Bigg( \sum_{j:\theta_j\in A_i}\pi_j -\pi_i^\star\Bigg)\frac{f_{\theta_i^\star}}{f^\star} + \sum_{j:\theta_j\in A_i} \pi_j\, \frac{f_{\theta_j}-f_{\theta_i^\star}}{f^\star} \Bigg\}. $$ Taylor expansion gives \begin{multline*} f_{\theta_j}(x)-f_{\theta_i^\star}(x)=(\theta_j-\theta_i^\star)^* D_1f_{\theta_i^\star}(x)+ \\ \frac{1}{2}\int_{0}^{1}(\theta_j-\theta_i^\star)^*D_2f_{\theta_i^\star+u(\theta_j-\theta_i^\star)}(x)\,(\theta_j-\theta_i^\star)\, 2(1-u)\,du. \end{multline*} Using Assumption A, we find that \begin{multline*} \left|\frac{f - f^\star}{f^\star}\right|\le \Bigg[ \sum_{j:\theta_j\in A_0}\pi_j +\sum_{i=1}^{q^\star}\Bigg\{ \Bigg| \sum_{j:\theta_j\in A_i}\pi_j -\pi_i^\star\Bigg| + \Bigg\| \sum_{j:\theta_j\in A_i}\pi_j(\theta_j-\theta_i^\star) \Bigg\| \\ \mbox{} + \frac{1}{2} \sum_{j:\theta_j\in A_i}\pi_j\|\theta_j-\theta_i^\star\|^2 \Bigg\}\Bigg] \,(H_0+H_1+H_2)\,d. \end{multline*} On the other hand, Theorem \ref{thm:localgeom} gives \begin{multline*} \left\|\frac{f - f^\star}{f^\star}\right\|_1\ge c^\star \Bigg[ \sum_{j:\theta_j\in A_0}\pi_j +\sum_{i=1}^{q^\star} \Bigg\{ \Bigg| \sum_{j:\theta_j\in A_i}\pi_j -\pi_i^\star\Bigg| \\ \mbox{} + \Bigg\| \sum_{j:\theta_j\in A_i}\pi_j(\theta_j-\theta_i^\star) \Bigg\| + \frac{1}{2} \sum_{j:\theta_j\in A_i}\pi_j\|\theta_j-\theta_i^\star\|^2 \Bigg\}\Bigg]. \end{multline*} The proof follows directly. \end{proof} \begin{cor} \label{cor:denvelope} $|d|\le D$ for all $d\in\mathcal{D}$, where $D=2S\in L^4(f^\star d\mu)$. \end{cor} \begin{proof} Using $\|f-f^\star\|_{\rm TV}\le 2h(f,f^\star)$ and $|\sqrt{x}-1|\le |x-1|$, we find $$ |d_f| = \frac{|\sqrt{f/f^\star}-1|}{h(f,f^\star)} \le \frac{|f/f^\star-1|}{\frac{1}{2}\|f/f^\star-1\|_1} \le 2S, $$ where we have used Lemma \ref{lem:senvelope}. \end{proof} Next, we prove that the Hellinger normalized densities $d_f$ can be approximated by chi-square normalized densities for small $h(f,f^\star)$. \begin{lem} \label{lem:chi2} For any $f\in\mathcal{M}$, we have $$ \left| \frac{\sqrt{f/f^\star}-1}{h(f,f^\star)} - \frac{f/f^\star-1}{\sqrt{\chi^2(f||f^\star)}} \right| \le \{4\|S\|_4^2 S+2S^2\}\,h(f,f^\star), $$ where we have defined the chi-square divergence $\chi^2(f||f^\star)=\|f/f^\star-1\|_2^2$. \end{lem} \begin{proof} Let us define the function $R$ as $$ \sqrt{\frac{f}{f^\star}}-1 = \frac{1}{2}\left\{ \frac{f-f^\star}{f^\star}+R \right\}. $$ Then we have \begin{multline*} \frac{\sqrt{f/f^\star}-1}{h(f,f^\star)} - \frac{f/f^\star-1}{\sqrt{\chi^2(f||f^\star)}} = \frac{f/f^\star-1+R}{\|f/f^\star-1+R\|_2}- \frac{f/f^\star-1}{\|f/f^\star-1\|_2} = \mbox{}\\ \frac{(f/f^\star-1+R)\{\|f/f^\star-1\|_2-\|f/f^\star-1+R\|_2\} +R\|f/f^\star-1+R\|_2 }{\|f/f^\star-1+R\|_2\,\|f/f^\star-1\|_2}, \end{multline*} so that by the reverse triangle inequality and Corollary \ref{cor:denvelope} $$ \left| \frac{\sqrt{f/f^\star}-1}{h(f,f^\star)} - \frac{f/f^\star-1}{\sqrt{\chi^2(f||f^\star)}}\right| \le \frac{2\|R\|_2S+|R|}{\|f/f^\star-1\|_2}. $$ Now note that for all $x\ge -1$ $$ -\frac{x^2}{2}\le -\frac{(\sqrt{1+x}-1)^2}{2} = \sqrt{1+x}-1-\frac{x}{2} \le 0. $$ Therefore, by Lemma \ref{lem:senvelope}, $$ |R| \le \left(\frac{f-f^\star}{f^\star}\right)^2 \le S^2\left\|\frac{f-f^\star}{f^\star}\right\|_1^2 \le S^2\left\|\frac{f-f^\star}{f^\star}\right\|_1 \left\|\frac{f-f^\star}{f^\star}\right\|_2. $$ The proof is easily completed using $\|f-f^\star\|_{\rm TV}\le 2h(f,f^\star)$. \end{proof} Finally, we need one further approximation step. \begin{lem} \label{lem:deriv} Let $q\in\mathbb{N}$ and $\alpha>0$. Then for every $f\in\mathcal{M}_q$ such that $h(f,f^\star)\le\alpha$, it is possible to choose coefficients $\eta_i\in\mathbb{R}$, $\beta_{i}\in\mathbb{R}^{d}$, $\rho_i\in M_+^d$ for $i=1,\ldots,q^{\star}$, and $\gamma_i\ge 0$, $\theta_i\in\Theta$ for $i=1,\ldots, q$, such that $\sum_{i=1}^{q^\star}\mathrm{rank}[\rho_i]\le q\wedge dq^\star$, $$ \displaylines{ \sum_{i=1}^{q^\star}|\eta_i|\le \frac{1}{c^\star}+\frac{1}{\sqrt{c^\star\alpha}}, \qquad \sum_{i=1}^{q^\star}\|\beta_i\|\le \frac{1}{c^\star}+\frac{2T}{\sqrt{c^\star\alpha}}, \cr \sum_{i=1}^{q^\star}\mathrm{Tr}[\rho_{i}]\le \frac{1}{c^\star}, \qquad \sum_{j=1}^q|\gamma_j|\le \frac{1}{\sqrt{c^\star\alpha}\wedge c^\star}, } $$ and $$ \left| \frac{f/f^\star-1}{\sqrt{\chi^2(f||f^\star)}} - \ell \right|\le \frac{d^{3/2}\sqrt{2}}{3(c^\star)^{5/4}}\, \{\|H_3\|_2\, S + H_3\}\,\alpha^{1/4}, $$ where we have defined $$ \ell = \sum_{i=1}^{q^\star}\left\{ \eta_i\frac{f_{\theta_i^\star}}{f^\star} +\beta_i^*\frac{D_1f_{\theta_i^\star}}{f^\star} +\mathrm{Tr}\Bigg[\rho_i \frac{D_2f_{\theta_i^\star}}{f^\star}\Bigg] \right\} + \sum_{j=1}^q\gamma_j\frac{f_{\theta_j}}{f^\star}. $$ \end{lem} \begin{proof} As $f\in\mathcal{M}_q$, we can write $f=\sum_{j=1}^q\pi_jf_{\theta_j}$. Note that by Theorem \ref{thm:localgeom} $$ h(f,f^\star) \ge \frac{c^\star}{4} \sum_{i=1}^{q^\star} \sum_{j:\theta_j\in A_i}\pi_j\|\theta_j-\theta_i^\star\|^2. $$ Therefore, $h(f,f^\star)\le\alpha$ implies $\pi_j\|\theta_j-\theta_i^\star\|^2\le 4\alpha/c^\star$ for $\theta_j\in A_i$. In particular, whenever $\theta_j\in A_i$, either $\pi_j\le 2\sqrt{\alpha/c^\star}$ or $\|\theta_j-\theta_i^\star\|^2\le 2\sqrt{\alpha/c^\star}$. Define $$ J=\bigcup_{i=1,\ldots,q^\star}\left\{j:\theta_j\in A_i,~ \|\theta_j-\theta_i^\star\|^2\le 2\sqrt{\alpha/c^\star}\right\}. $$ Taylor expansion gives $$ f_{\theta_j}(x)-f_{\theta_i^\star}(x)=(\theta_j-\theta_i^\star)^* D_1f_{\theta_i^\star}(x) +\frac{1}{2}(\theta_j-\theta_i^\star)^*D_2f_{\theta_i^\star}(x) \,(\theta_j-\theta_i^\star) + R_{ji}(x), $$ where $|R_{ji}| \le \frac{1}{6}d^{3/2}\|\theta_j-\theta_i^\star\|^3H_3$. We can therefore write $$ \frac{f-f^\star}{f^\star} = L + \sum_{i=1}^{q^\star}\sum_{j\in J:\theta_j\in A_i} \pi_j R_{ji}, $$ where we have defined \begin{multline*} L = \sum_{i=1}^{q^\star}\Bigg\{ \Bigg(\sum_{j\in J:\theta_j\in A_i}\pi_j-\pi_i^\star\Bigg) \frac{f_{\theta_i^\star}}{f^\star} + \sum_{j\in J:\theta_j\in A_i}\pi_j(\theta_j-\theta_i^\star)^*\, \frac{D_1f_{\theta_i^\star}}{f^\star} \\ \mbox{} + \frac{1}{2} \sum_{j\in J:\theta_j\in A_i}\pi_j(\theta_j-\theta_i^\star)^*\, \frac{D_2f_{\theta_i^\star}}{f^\star}(\theta_j-\theta_i^\star)\Bigg\} + \sum_{j\not\in J}\pi_j\,\frac{f_{\theta_j}}{f^\star}. \end{multline*} Now note that \begin{align*} \left| \frac{f/f^\star-1}{\sqrt{\chi^2(f||f^\star)}} - \frac{L}{\|L\|_2} \right| &\le \frac{|f/f^\star-1|}{\|f/f^\star-1\|_2} \frac{\|f/f^\star-1-L\|_2}{\|L\|_2} + \frac{|f/f^\star-1-L|}{\|L\|_2} \\ &\le \frac{\|f/f^\star-1-L\|_2\,S+|f/f^\star-1-L|}{\|L\|_2}, \end{align*} where we have used Lemma \ref{lem:senvelope}. By Theorem \ref{thm:localgeom}, we obtain $$ \|L\|_2 \ge \|L\|_1 \ge \frac{c^\star}{2}\sum_{i=1}^{q^\star} \sum_{j\in J:\theta_j\in A_i}\pi_j\|\theta_j-\theta_i^\star\|^2. $$ Therefore, we can estimate $$ \frac{|f/f^\star-1-L|}{\|L\|_2} \le \frac{d^{3/2}H_3}{3c^\star} \frac{\sum_{i=1}^{q^\star}\sum_{j\in J:\theta_j\in A_i} \pi_j\|\theta_j-\theta_i^\star\|^3}{ \sum_{i=1}^{q^\star} \sum_{j\in J:\theta_j\in A_i}\pi_j\|\theta_j-\theta_i^\star\|^2} \le \left(\frac{4\alpha}{c^\star}\right)^{1/4} \frac{d^{3/2}H_3}{3c^\star} $$ where we have used the definition of $J$. Setting $\ell=L/\|L\|_2$, we obtain $$ \left| \frac{f/f^\star-1}{\sqrt{\chi^2(f||f^\star)}} - \ell \right| \le \frac{d^{3/2}\sqrt{2}}{3(c^\star)^{5/4}}\, \{\|H_3\|_2\, S + H_3\}\,\alpha^{1/4}. $$ It remains to show that for our choice of $\ell=L/\|L\|_2$, the coefficients $\eta,\beta,\rho,\gamma$ in the statement of the lemma satisfy the desired bounds. These coefficients are $$ \displaylines{ \eta_i=\frac{1}{\|L\|_2}\Bigg( \sum_{j\in J:\theta_j\in A_i}\pi_j-\pi_i^\star\Bigg),\qquad \beta_i=\frac{1}{\|L\|_2}\sum_{j\in J:\theta_j\in A_i} \pi_j(\theta_j-\theta_i^\star), \cr \rho_{i}=\frac{1}{2\|L\|_2} \sum_{j\in J:\theta_j\in A_i} \pi_j(\theta_j-\theta_i^\star) (\theta_j-\theta_i^\star)^*, \qquad \gamma_j=\frac{\pi_j\mathbf{1}_{j\not\in J}}{\|L\|_2}. } $$ Clearly $\mathrm{rank}[\rho_i]\le \#\{j:\theta_j\in A_i\}\wedge d$, so $\sum_{i=1}^{q^\star}\mathrm{rank}[\rho_i]\le q\wedge dq^\star$. Moreover, \begin{multline*} \|L\|_2 \ge c^\star\Bigg[\sum_{j:\theta_j\in A_0}\pi_j + \sum_{i=1}^{q^\star} \Bigg\{ \Bigg|\sum_{j:\theta_j\in A_i}\pi_j-\pi_i^\star\Bigg| \\ +\Bigg\| \sum_{j:\theta_j\in A_i}\pi_j(\theta_j-\theta_i^\star)\Bigg\| + \frac{1}{2} \sum_{j:\theta_j\in A_i}\pi_j\|\theta_j-\theta_i^\star\|^2\Bigg\} \Bigg] \end{multline*} by Theorem \ref{thm:localgeom}. It follows that $\sum_{i=1}^{q^\star}\mathrm{Tr}[\rho_{i}]\le 1/c^\star$. Now note that for $j\not\in J$ such that $\theta_j\in A_i$, we have $\|\theta_j-\theta_i^\star\|^2>2\sqrt{\alpha/c^\star}$ by construction. Therefore $$ \|L\|_2 \ge c^\star\Bigg[ \sum_{j\not\in J:\theta_j\in A_0}\pi_j + \frac{1}{2}\sum_{i=1}^{q^\star} \sum_{j\not\in J:\theta_j\in A_i} \pi_j\|\theta_j-\theta_i^\star\|^2 \Bigg] \ge ( \sqrt{c^\star\alpha} \wedge c^\star )\,\sum_{j\not\in J}\pi_j. $$ It follows that $\sum_{j=1}^q|\gamma_j|\le 1/ ( \sqrt{c^\star\alpha} \wedge c^\star )$. Next, we note that $$ \sum_{i=1}^{q^\star} \Bigg|\sum_{j\in J:\theta_j\in A_i}\pi_j-\pi_i^\star\Bigg| \le \sum_{i=1}^{q^\star} \Bigg|\sum_{j:\theta_j\in A_i}\pi_j-\pi_i^\star\Bigg| + \sum_{j\not\in J:\theta_j\not\in A_0}\pi_j. $$ Therefore $\sum_{i=1}^{q^\star}|\eta_i|\le 1/c^\star+1/\sqrt{c^\star\alpha}$. Finally, note that $$ \sum_{i=1}^{q^\star} \Bigg\|\sum_{j\in J:\theta_j\in A_i} \pi_j(\theta_j-\theta_i^\star)\Bigg\| \le \sum_{i=1}^{q^\star} \Bigg\|\sum_{j:\theta_j\in A_i} \pi_j(\theta_j-\theta_i^\star)\Bigg\| + 2T\sum_{j\not\in J:\theta_j\not\in A_0}\pi_j. $$ Therefore $\sum_{i=1}^{q^\star}\|\beta_i\|\le 1/c^\star+2T/\sqrt{c^\star\alpha}$. The proof is complete. \end{proof} We can now complete the proof of Theorem \ref{thm:mainmixtures}. \begin{proof}[Proof of Theorem \ref{thm:mainmixtures}] Let $\alpha>0$ be a constant to be chosen later on, and $$ \mathcal{D}_{q,\alpha}=\{d_f:f\in\mathcal{M}_q,~ f\ne f^\star,~ h(f,f^\star)\le\alpha\}. $$ Then clearly $$ \mathcal{N}(\mathcal{D}_q,\delta) \le \mathcal{N}(\mathcal{D}_{q,\alpha},\delta) + \mathcal{N}(\mathcal{D}_q\backslash\mathcal{D}_{q,\alpha},\delta). $$ We will estimate each term separately. \textbf{Step 1} (\emph{the first term}). Define $$ \mathbb{M}_{q}= \{(m_{1},\ldots,m_{q^\star})\in \mathbb{Z}_+^{q^\star}:m_{1}+\cdots+m_{q^\star}=q \wedge dq^\star\}. $$ For every $m\in\mathbb{M}_q$, we define the family of functions \begin{multline*} \mathcal{L}_{q,m,\alpha}= \Bigg\{ \sum_{i=1}^{q^\star}\left\{ \eta_i\frac{f_{\theta_i^\star}}{f^\star} +\beta_i^*\frac{D_1f_{\theta_i^\star}}{f^\star} +\sum_{j=1}^{m_{i}}\rho_{ij}^*\frac{D_2f_{\theta_i^\star}}{f^\star}\rho_{ij} \right\} + \sum_{j=1}^{q}\gamma_j\frac{f_{\theta_j}}{f^{\star}}:\mbox{}\\ (\eta,\beta,\rho,\gamma,\theta)\in\mathfrak{I}_{q,m,\alpha}\Bigg\}, \end{multline*} where \begin{multline*} \mathfrak{I}_{q,m,\alpha}=\Bigg\{(\eta,\beta,\rho,\gamma,\theta)\in \mathbb{R}^{q^\star}\times (\mathbb{R}^d)^{q^\star}\times (\mathbb{R}^d)^{m_{1}}\times\cdots\times (\mathbb{R}^d)^{m_{q^\star}} \times \mathbb{R}^q\times\Theta^q: \mbox{}\\ \sum_{i=1}^{q^\star}|\eta_i|\le \frac{1}{c^\star}+\frac{1}{\sqrt{c^\star\alpha}}, \qquad \sum_{i=1}^{q^\star}\|\beta_i\|\le \frac{1}{c^\star}+\frac{2T}{\sqrt{c^\star\alpha}}, \mbox{}\\ \sum_{i=1}^{q^\star}\sum_{j=1}^{m_{i}}\|\rho_{ij}\|^2\le \frac{1}{c^\star}, \qquad \sum_{j=1}^q|\gamma_j|\le \frac{1}{\sqrt{c^\star\alpha}\wedge c^\star} \Bigg\}. \end{multline*} Define the family of functions $$ \mathcal{L}_{q,\alpha}= \bigcup_{m\in\mathbb{M}_{q}}\mathcal{L}_{q,m,\alpha} $$ From Lemmas \ref{lem:chi2} and \ref{lem:deriv}, we find that for any function $d\in\mathcal{D}_{q,\alpha}$, there exists a function $\ell\in \mathcal{L}_{q,\alpha}$ such that (here we use that $h(f,f^\star)\le\sqrt{2}$ for any $f$) $$ |d-\ell| \le \{4\|S\|_4^2 S+2S^2\}\,(\alpha\wedge\sqrt{2}) + \frac{d^{3/2}\sqrt{2}}{3(c^\star)^{5/4}}\, \{\|H_3\|_2\, S + H_3\}\,\alpha^{1/4}. $$ Using $\alpha\wedge\sqrt{2}\le 2^{3/8}\alpha^{1/4}$ for all $\alpha>0$, we can estimate $$ |d-\ell| \le \alpha^{1/4}\,U,\qquad U=\left( \frac{1+\|H_3\|_2}{(c^\star)^{5/4}}+ 8\|S\|_4^2 + 4\right)d^{3/2}\,\{S+S^2+H_3\}, $$ where $U\in L^2(f^\star d\mu)$ by Assumption A. Now note that if $m_1\le\ell\le m_2$ for some functions $m_1,m_2$ with $\|m_2-m_1\|_2\le\varepsilon$, then $m_1-\alpha^{1/4}\,U\le d\le m_2+\alpha^{1/4}\,U$ with $\|(m_2+\alpha^{1/4}\,U) -(m_1-\alpha^{1/4}\,U)\|_2\le \varepsilon+2\alpha^{1/4}\|U\|_2$. Therefore $$ \mathcal{N}(\mathcal{D}_{q,\alpha}, \varepsilon+2\alpha^{1/4}\|U\|_2) \le \mathcal{N}(\mathcal{L}_{q,\alpha},\varepsilon) \le \sum_{m\in\mathbb{M}_{q}}\mathcal{N}(\mathcal{L}_{q,m,\alpha},\varepsilon) \quad \mbox{for }\varepsilon>0. $$ Of course, we will ultimately choose $\varepsilon,\alpha$ such that $\varepsilon+2\alpha^{1/4}\|U\|_2=\delta$. We proceed to estimate the bracketing number $\mathcal{N}(\mathcal{L}_{q,m,\alpha},\varepsilon)$. To this end, let $\ell,\ell'\in\mathcal{L}_{q,m,\alpha}$, where $\ell$ is defined by the parameters $(\eta,\beta,\rho,\gamma,\theta)\in\mathfrak{I}_{q,m,\alpha}$ and $\ell'$ is defined by the parameters $(\eta',\beta',\rho',\gamma',\theta')\in\mathfrak{I}_{q,m,\alpha}$. Note that $$ \sum_{i=1}^{q^\star}\sum_{j=1}^{m_i} \Bigg| \rho_{ij}^*\frac{D_2f_{\theta_i^\star}}{f^\star}\rho_{ij}- (\rho_{ij}')^*\frac{D_2f_{\theta_i^\star}}{f^\star}\rho_{ij}' \Bigg| \le \frac{2d}{\sqrt{c^\star}}\,H_2 \sum_{i=1}^{q^\star}\sum_{j=1}^{m_i} \,\|\rho_{ij}-\rho_{ij}'\|. $$ We can therefore estimate \begin{multline*} |\ell-\ell'| \le H_0\sum_{i=1}^{q^\star}|\eta_i-\eta_i'| + H_1\sqrt{d}\sum_{i=1}^{q^\star}\|\beta_i-\beta_i'\| + H_0\sum_{j=1}^q|\gamma_j-\gamma_j'| + \\ \frac{\sqrt{d}}{\sqrt{c^\star\alpha}\wedge c^\star} \,H_1 \max_{j=1,\ldots,q}\|\theta_j-\theta_j'\| + \frac{2d\sqrt{dq^\star}}{\sqrt{c^\star}}\,H_2\, \Bigg[ \sum_{i=1}^{q^\star}\sum_{j=1}^{m_i} \,\|\rho_{ij}-\rho_{ij}'\|^2\Bigg]^{1/2}. \end{multline*} where we have used that $|f_{\theta}-f_{\theta'}|/f^\star \le \|\theta-\theta'\|\,H_1 \sqrt{d}$ by Taylor expansion. Therefore, writing $V=(H_0+H_1+H_2)\,d\sqrt{dq^\star}$, we have $$ |\ell-\ell'| \le V\, \tnorm{(\eta,\beta,\rho,\gamma,\theta)- (\eta',\beta',\rho',\gamma',\theta')}_{q,m,\alpha}, $$ where $\tnorm{\cdot}_{q,m,\alpha}$ is the norm on $\mathbb{R}^{(1+d)q^\star+ d(q\wedge dq^\star)+(1+d)q}$ defined by \begin{multline*} \tnorm{(\eta,\beta,\rho,\gamma,\theta)}_{q,m,\alpha} = \sum_{i=1}^{q^\star}|\eta_i| + \sum_{i=1}^{q^\star}\|\beta_i\| + \sum_{j=1}^q|\gamma_j| \\ + \frac{1}{\sqrt{c^\star\alpha}\wedge c^\star} \max_{j=1,\ldots,q}\|\theta_j\| + \frac{2}{\sqrt{c^\star}}\, \Bigg[ \sum_{i=1}^{q^\star}\sum_{j=1}^{m_i} \,\|\rho_{ij}\|^2\Bigg]^{1/2}. \end{multline*} Note that if $\tnorm{(\eta,\beta,\rho,\gamma,\theta)- (\eta',\beta',\rho',\gamma',\theta')}_{q,m,\alpha}\le \varepsilon'$, then we obtain a bracket $\ell'-\varepsilon'V\le \ell \le \ell'+\varepsilon'V$ of size $\|(\ell'+\varepsilon'V)-(\ell'-\varepsilon'V)\|_2= 2\varepsilon'\|V\|_2$. Therefore, if we denote by $N(\mathfrak{I}_{q,m,\alpha},\tnorm{\cdot}_{q,m,\alpha},\varepsilon')$ the cardinality of the largest packing of $\mathfrak{I}_{q,m,\alpha}$ by $\varepsilon'$-separated points with respect to the $\tnorm{\cdot}_{q,m,\alpha}$-norm, then $$ \mathcal{N}(\mathcal{L}_{q,m,\alpha},\varepsilon) \le N(\mathfrak{I}_{q,m,\alpha},\tnorm{\cdot}_{q,m,\alpha}, \varepsilon/2\|V\|_2)\quad \mbox{for }\varepsilon>0. $$ But note that, by construction, $\mathfrak{I}_{q,m,\alpha}$ is included in a $\tnorm{\cdot}_{q,m,\alpha}$-ball of radius not exceeding $(6+3T)/(\sqrt{c^\star\alpha}\wedge c^\star)$. Therefore, using the standard fact that the packing number of the $r$-ball $B(r)=\{x\in B:\tnorm{x}\le r\}$ in any $n$-dimensional normed space $(B,\tnorm{\cdot})$ satisfies $N(B(r),\tnorm{\cdot},\varepsilon)\le (\frac{2r+\varepsilon}{\varepsilon})^n$, we can estimate $$ \mathcal{N}(\mathcal{L}_{q,m,\alpha},\varepsilon) \le \left(\frac{4\|V\|_2(6+3T)/(\sqrt{c^\star\alpha}\wedge c^\star) +\varepsilon}{\varepsilon} \right)^{(1+d)q^\star+ d(q\wedge dq^\star)+(1+d)q}. $$ In particular, if $\varepsilon\le 1$ and $\alpha\le c^\star$, then $$ \mathcal{N}(\mathcal{L}_{q,m,\alpha},\varepsilon) \le \left(\frac{(24+12T)\|V\|_2/\sqrt{c^\star} +\sqrt{c^\star}}{\varepsilon\sqrt{\alpha}} \right)^{3(d+1)q}. $$ Finally, note that the cardinality of $\mathbb{M}_q$ can be estimated as $$ \#\mathbb{M}_q = {q^\star + q\wedge dq^\star -1 \choose q\wedge dq^\star} \le e^{q\wedge dq^\star} \left(\frac{q^\star+q\wedge dq^\star -1}{q\wedge dq^\star} \right)^{q\wedge dq^\star} \le 2^{3q}, $$ where we have used that $q\ge q^\star$. We therefore obtain \begin{align*} \mathcal{N}(\mathcal{D}_{q,\alpha},\delta) &\le \sum_{m\in\mathbb{M}_{q}} \mathcal{N}(\mathcal{L}_{q,m,\alpha}, \delta-2\alpha^{1/4}\|U\|_2) \\ &\le \left(\frac{24(2+T)\|V\|_2/\sqrt{c^\star} +\sqrt{c^\star}}{(\delta-2\alpha^{1/4}\|U\|_2)\sqrt{\alpha}} \right)^{3(d+1)q} \end{align*} whenever $\delta\le 1$ and $\alpha\le (\delta/2\|U\|_2)^4\wedge c^\star$. \textbf{Step 2} (\emph{the second term}). For $f,f'\in\mathcal{M}_q$ with $h(f,f^\star)>\alpha$ and $h(f',f^\star)>\alpha$, \begin{align*} |d_f-d_f'| &= \frac{|(\sqrt{f/f^\star}-1)\|\sqrt{f'/f^\star}-1\|_2- (\sqrt{f'/f^\star}-1)\|\sqrt{f/f^\star}-1\|_2| }{h(f,f^\star)\,h(f',f^\star)} \\ &\le \frac{ \|\sqrt{f'/f^\star}-\sqrt{f/f^\star}\|_2 |\sqrt{f/f^\star}-1| + \sqrt{2}\,|\sqrt{f/f^\star}-\sqrt{f'/f^\star}| }{\alpha^2}, \end{align*} where we have used that $h(f,f^\star)\le\sqrt{2}$ for any $f$. Now note that $$ \big|\sqrt{a}-\sqrt{b}\big|^2 \le \big|\sqrt{a}-\sqrt{b}\big|\left(\sqrt{a}+\sqrt{b}\right) =|a-b| $$ for any $a,b\ge 0$. We can therefore estimate $$ |d_f-d_f'| \le \frac{ \|(f-f')/f^\star\|_1^{1/2} (\sqrt{H_0}+1) + \sqrt{2}\,|(f-f')/f^\star|^{1/2}}{\alpha^2}, $$ where we have used that $|\sqrt{f/f^\star}-1|\le \sqrt{H_0}+1$ for any $f\in\mathcal{M}$. Now note that if we write $f=\sum_{i=1}^q\pi_if_{\theta_i}$ and $f'=\sum_{i=1}^q\pi_i'f_{\theta_i'}$, then we can estimate $$ \left|\frac{f-f'}{f^\star}\right| \le H_0\sum_{i=1}^q|\pi_i-\pi_i'| + H_1\sqrt{d}\max_{i=1,\ldots,q}\|\theta_i-\theta_i'\|. $$ Defining $$ W = (\sqrt{H_0}+1) \|H_0+H_1\sqrt{d}\|_1^{1/2} + \sqrt{2}\,(H_0+H_1\sqrt{d})^{1/2}, $$ we obtain $$ |d_f-d_f'| \le \frac{W}{\alpha^2}\,\tnorm{(\pi,\theta)-(\pi',\theta')}_q^{1/2}, \qquad \tnorm{(\pi,\theta)}_q=\sum_{i=1}^q|\pi_i| + \max_{i=1,\ldots,q}\|\theta_i\| $$ (clearly $\tnorm{\cdot}_q$ defines a norm on $\mathbb{R}^{(d+1)q}$). Now note that if $\tnorm{(\pi,\theta)-(\pi',\theta')}_q\le\varepsilon$, then we obtain a bracket $d_f'-\varepsilon^{1/2}W/\alpha^2\le d_f\le d_f'+\varepsilon^{1/2}W/\alpha^2$ of size $\|(d_f'+\varepsilon^{1/2}W/\alpha^2)-(d_f'-\varepsilon^{1/2}W/\alpha^2)\|_2 = 2\varepsilon^{1/2}\|W\|_2/\alpha^2$. Therefore $$ \mathcal{N}(\mathcal{D}_q\backslash\mathcal{D}_{q,\alpha},\delta) \le N(\Delta_q\times\Theta^q, \tnorm{\cdot}_q,\alpha^4\delta^2/4\|W\|_2^2), $$ where we have defined the simplex $\Delta_q=\{\pi\in\mathbb{R}^q_+:\sum_{i=1}^q\pi_i=1\}$. We can now estimate the quantity on the right hand side of this expression as before, giving $$ \mathcal{N}(\mathcal{D}_q\backslash\mathcal{D}_{q,\alpha},\delta) \le \left( \frac{8(1+T)\|W\|_2^2 +(c^\star)^4}{\alpha^4\delta^2} \right)^{(d+1)q} $$ for $\delta\le 1$ and $\alpha\le c^\star$. \textbf{End of proof.} Choose $\alpha=(\delta/4\|U\|_2)^4$. Collecting the various estimates above, we find that for $\delta\le 1\wedge 4(c^\star)^{1/4}$ (as $\|U\|_2\ge\|S\|_1\ge 1$ by Lemma \ref{lem:senvelope}) \begin{align*} \mathcal{N}(\mathcal{D}_{q},\delta) &\le \left(\frac{768(2+T)\|U\|_2^2\|V\|_2/\sqrt{c^\star} +32\|U\|_2^2\sqrt{c^\star}}{\delta^3} \right)^{3(d+1)q} \\ &\qquad\quad\mbox{}+ \left( \frac{4^{18}(1+T)\|U\|_2^{16}\|W\|_2^2 + 4^{16}\|U\|_2^{16}(c^\star)^4}{\delta^{18}} \right)^{(d+1)q} \\ &\le \left( \frac{ c_0^\star\,(T\vee 1)^{1/6}\, (\|U\|_2\vee\|V\|_2\vee\|W\|_2) }{\delta} \right)^{18(d+1)q} \end{align*} where $c_0^\star = 12(c^\star)^{-1/12} + 2(c^\star)^{1/12} + 4(c^\star)^{4/18} + 8$. It follows that $$ \mathcal{N}(\mathcal{D}_{q},\delta) \le \left( \frac{ C^\star (T\vee 1)^{1/6} (\|H_0\|_4^4\vee\|H_1\|_4^4\vee\|H_2\|_4^4\vee\|H_3\|_2^2) }{\delta} \right)^{18(d+1)q} $$ for all $\delta\le\delta^\star$, where $C^\star$ and $\delta^\star$ are constants that depend only on $c^\star$, $d$, and $q^\star$. This establishes the estimate given in the statement of the Theorem. The proof of the second half of the Theorem follows from Corollary \ref{cor:denvelope} and $\|H_0\|_4\ge 1$. \end{proof} \subsection{Proof of Theorem \ref{thm:mainconsist}} \label{sec:proofmainconsist} The proof of Theorem \ref{thm:mainconsist} is based on Theorem \ref{thm:upperlil} and the following result. \begin{prop} \label{prop:liklln} Let $\mathcal{M}^n$ for $n\ge 1$ be a family of strictly positive probability densities with respect to a reference measure $\mu$ such that $\mathcal{M}^n\subseteq\mathcal{M}^{n+1}$ for all $n$. Define $\mathcal{M}=\bigcup_n\mathcal{M}^n$, and let $f^\star$ be another probability density with respect to $\mu$ such that $f^\star\not\in\mathop{\mathrm{cl}}\mathcal{M}$, where $\mathop{\mathrm{cl}}\mathcal{M}$ denotes the $L^1(d\mu)$-closure of $\mathcal{M}$. Let $\mathcal{H}^n = \{\sqrt{f/f^\star}:f\in\mathcal{M}^n\}$, and suppose there exist $K(n)\ge 1$ and $p\ge 1$ so that $$ \mathcal{N}(\mathcal{H}^n,\delta) \le\left( \frac{K(n)}{\delta} \right)^p $$ for all $\delta\le 1$ and $n\ge 1$, where $\mathcal{N}(\mathcal{H}^n,\delta)$ is the minimal number of brackets of $L^2(f^\star d\mu)$-width $\delta$ needed to cover $\mathcal{H}^n$. Let $(X_i)_{i\in\mathbb{N}}$ be i.i.d.\ with distribution $f^\star d\mu$. If in addition $\log K(n)=o(n)$, then we have $$ \limsup_{n\to\infty}\sup_{f\in\mathcal{M}^n} \frac{1}{n}\sum_{j=1}^n\log\left( \frac{f(X_j)}{f^\star(X_j)} \right)<0\quad\mbox{a.s.} $$ \end{prop} \begin{proof} As in the proof of Theorem \ref{thm:deviationbound}, we have $$ \frac{1}{n}\sum_{j=1}^n\log\left(\frac{f(X_j)}{f^\star(X_j)}\right) \le 4n^{-1/2} \nu_n(\log(\{\bar f/f^\star\}^{1/2})) - 2 D(f^\star||\bar f). $$ The following claim will be proved below: $$ \lim_{n\to\infty} \sup_{f\in\mathcal{M}^n}n^{-1/2} \nu_n(\log(\{\bar f/f^\star\}^{1/2})) =0\quad\mbox{a.s.} $$ Using the claim, the proof is easily completed: indeed, we then have $$ \limsup_{n\to\infty}\sup_{f\in\mathcal{M}^n} \frac{1}{n}\sum_{j=1}^n\log\left( \frac{f(X_j)}{f^\star(X_j)} \right) \le -2\inf_{f\in\mathcal{M}} D(f^\star||\bar f)<0\quad\mbox{a.s.}, $$ where the last inequality follows from Pinsker's inequality and $f^\star\not\in\mathop{\mathrm{cl}}\mathcal{M}$. It therefore remains to prove the claim. To this end we apply \cite{vdG00}, Theorem 5.11 as in the proof of \cite{vdG00}, Theorem 7.4 (cf.\ Theorem \ref{thm:deviationbound} above), which yields $$ \mathbf{P}\left[ \sup_{f\in\mathcal{M}^n}|n^{-1/2} \nu_n(\log(\{\bar f/f^\star\}^{1/2}))| \ge\alpha \right]\le C\,e^{-n\alpha^2/C} $$ for every $\alpha>0$ such that $C\sqrt{p}\,(1+\sqrt{\log K(n)})\le\alpha\sqrt{n}\le 32\sqrt{n}$ and $n\ge 1$, where $C$ is a universal constant. As $\log K(n)=o(n)$, we have $$ \sum_{n\ge 1} \mathbf{P}\left[ \sup_{f\in\mathcal{M}^n}|n^{-1/2} \nu_n(\log(\{\bar f/f^\star\}^{1/2}))| \ge\alpha \right]<\infty $$ for all $0<\alpha\le 32$, so the claim follows from the Borel-Cantelli lemma. \end{proof} We can now complete the proof of Theorem \ref{thm:mainconsist}. \begin{proof}[Proof of Theorem \ref{thm:mainconsist}] By Theorem \ref{thm:upperlil} and easy manipulations, $\mathbf{P}^\star$-a.s. \begin{align*} &\limsup_{n\to\infty} \sup_{q>q^\star} \frac{1}{\mathrm{pen}(n,q)-\mathrm{pen}(n,q^\star)} \left\{ \sup_{f\in\mathcal{M}_q^n}\ell_n(f)- \sup_{f\in\mathcal{M}_{q^\star}^n}\ell_n(f) \right\} \\ &\mbox{}\le \lim_{n\to\infty} \sup_{q>q^\star} \frac{\eta(q)\{\log K(2n)\vee \log\log n\}}{\mathrm{pen}(n,q)- \mathrm{pen}(n,q^\star)} \times\mbox{} \\ & \quad\limsup_{n\to\infty} \frac{1}{\log K(2n)\vee \log\log n} \sup_{q>q^\star} \frac{1}{\eta(q)} \left\{ \sup_{f\in\mathcal{M}_q^n}\ell_n(f)- \sup_{f\in\mathcal{M}_{q^\star}^n}\ell_n(f) \right\} =0. \end{align*} Therefore, $\mathbf{P}^\star$-a.s.\ eventually as $n\to\infty$ $$ \sup_{f\in\mathcal{M}_q^n}\ell_n(f) - \mathrm{pen}(n,q) < \sup_{f\in\mathcal{M}_{q^\star}^n}\ell_n(f) - \mathrm{pen}(n,q^\star) $$ for all $q>q^\star$. It follows that $\limsup_{n\to\infty}\hat q_n\le q^\star$ $\mathbf{P}^\star$-a.s., that is, the penalized likelihood order estimator does not asymptotically overestimate the order. On the other hand, we note that for every $q<q^\star$ $$ \limsup_{n\to\infty}\frac{1}{n}\left\{ \sup_{f\in\mathcal{M}_q^n}\ell_n(f)- \sup_{f\in\mathcal{M}_{q^\star}^n}\ell_n(f) \right\} \le \limsup_{n\to\infty} \sup_{f\in\mathcal{M}_q^n} \frac{1}{n}\sum_{j=1}^n \log\left( \frac{f(X_j)}{f^\star(X_j)}\right) $$ which is strictly negative $\mathbf{P}^\star$-a.s.\ by Proposition \ref{prop:liklln}, where we have used that $\log K(n)=o(n)$ and that $\mathcal{N}(\mathcal{H}_q^n(2),\delta)\le \mathcal{N}(\mathcal{H}_{q^\star}^n(2),\delta) \le (2K(n)/\delta)^{\eta(q^\star)}$ for all $\delta\le 2$ and $n$ sufficiently large. As $\mathrm{pen}(n,q)/n\to 0$ as $n\to\infty$ for $q<q^\star$ $$ \limsup_{n\to\infty}\max_{q<q^\star}\frac{1}{n}\left\{ \sup_{f\in\mathcal{M}_q^n}\ell_n(f)-\mathrm{pen}(n,q) -\sup_{f\in\mathcal{M}_{q^\star}^n}\ell_n(f)+ \mathrm{pen}(n,q^\star) \right\} < 0 $$ $\mathbf{P}^\star$-a.s. In particular, we find that $\mathbf{P}^\star$-a.s.\ eventually as $n\to\infty$ $$ \sup_{f\in\mathcal{M}_q^n}\ell_n(f) - \mathrm{pen}(n,q) < \sup_{f\in\mathcal{M}_{q^\star}^n}\ell_n(f) - \mathrm{pen}(n,q^\star) $$ for all $q<q^\star$. It follows that $\liminf_{n\to\infty}\hat q_n\ge q^\star$ $\mathbf{P}^\star$-a.s., that is, the penalized likelihood order estimator does not asymptotically underestimate the order. \end{proof} Finally, let us prove Corollary \ref{cor:inconsist}. \begin{proof}[Proof of Corollary \ref{cor:inconsist}] It is shown in the proof of Corollary \ref{cor:lowerlil} that $$ \Gamma:= \sup_{g\in L^2_0(f^\star d\mu)}\Bigg\{ \sup_{d\in\mathcal{\bar D}_q^c}(\langle d,g\rangle)_+^2 -\sup_{d\in\mathcal{\bar D}_{q^\star}}(\langle d,g\rangle)_+^2 \Bigg\}>0. $$ By Theorem \ref{thm:lowerlil}, we have \begin{multline*} \limsup_{n\to\infty} \frac{1}{\mathrm{pen}(n,q)-\mathrm{pen}(n,q^\star)} \Bigg\{\sup_{f\in\mathcal{M}_q}\ell_n(f)- \sup_{f\in\mathcal{M}_{q^\star}}\ell_n(f)\Bigg\} \ge \mbox{}\\ \frac{1}{C\{\eta(q)-\eta(q^\star)\}} \sup_{g\in L^2_0(f^\star d\mu)}\Bigg\{ \sup_{d\in\mathcal{\bar D}_q^c}(\langle d,g\rangle)_+^2 -\sup_{d\in\mathcal{\bar D}_{q^\star}}(\langle d,g\rangle)_+^2 \Bigg\} \quad\mathbf{P}^\star\mbox{-a.s.} \end{multline*} Therefore, choosing $C<\Gamma/\{\eta(q)-\eta(q^\star)\}$, we find that $$ \sup_{f\in\mathcal{M}_q}\ell_n(f)-\mathrm{pen}(n,q) > \sup_{f\in\mathcal{M}_{q^\star}}\ell_n(f) -\mathrm{pen}(n,q^\star) $$ infinitely often $\mathbf{P}^\star$-a.s., so that $\hat q_n\ne q^\star$ infinitely often $\mathbf{P}^\star$-a.s. \end{proof} \subsection{Proof of Proposition \ref{prop:mixconsist}} \label{sec:proofmixconsist} The proofs of the consistency results in Propositions \ref{prop:mixconsist} and \ref{prop:mixconsist2} follow almost immediately from Theorem \ref{thm:mainconsist}, Corollary \ref{cor:localmixtures}, and Example \ref{ex:gaussian}. The main difficulty is to establish the condition $\mathcal{\bar D}_q^c\backslash\mathcal{\bar D}_{q^\star}\ne \varnothing$ of Corollary \ref{cor:inconsist}, which is needed to prove the inconsistency part of Proposition \ref{prop:mixconsist}. To this end, we will need the following lemma characterizing $\mathcal{\bar D}_{q^\star}$ (here we adopt the same notations as in section \ref{sec:mixtures}). \begin{lem} \label{lem:dqstar} Suppose that Assumption A holds. Then we have $$ \mathcal{\bar D}_{q^\star} = \Bigg\{\frac{L}{\|L\|_2}: L=\sum_{i=1}^{q^\star} \Bigg\{\eta_i\,\frac{f_{\theta_i^\star}}{f^\star} +\beta_i^*\,\frac{D_1f_{\theta_i^\star}}{f^\star} \Bigg\}, ~ \eta_i\in\mathbb{R}, ~ \beta_i\in\mathbb{R}^{d}, ~ \sum_{i=1}^{q^\star}\eta_i=0 \Bigg\}. $$ \end{lem} \begin{proof} Let $(f_n)_{n\ge 1}\subset\mathcal{M}_{q^\star}$ be such that $h(f_n,f^\star)\to 0$ and $d_{f_n}\to d_0\in\mathcal{\bar D}_{q^\star}$. By Theorem \ref{thm:localgeom}, we may assume without loss of generality that $f_n=\sum_{i=1}^{q^\star}\pi_i^nf_{\theta_i^n}$ with $\theta_i^n\to\theta_i^\star$ and $\pi_i^n\to\pi_i^\star$ for every $i=1,\ldots,q^\star$. Taylor expansion gives $$ \frac{f_n-f^\star}{f^\star} = L_n + R_n,\qquad \quad |R_n| \le \frac{d}{2}\,H_2\sum_{i=1}^{q^\star} \pi_i^n\|\theta_i^n-\theta_i^\star\|^2, $$ where $$ L_n = \sum_{i=1}^{q^\star}\left\{ (\pi_i^n-\pi_i^\star)\,\frac{f_{\theta_i^\star}}{f^\star} +\pi_i^n(\theta_i^n-\theta_i^\star)^*\, \frac{D_1f_{\theta_i^\star}}{f^\star} \right\}. $$ Proceeding as in Lemmas \ref{lem:chi2} and \ref{lem:deriv}, we can estimate $$ \left\|d_{f_n}- \frac{L_n}{\|L_n\|_2}\right\|_2 \le 2\|S\|_4^2\{2\|S\|_2+1\}\,h(f_n,f^\star)+ \{\|S\|_2+1\}\,\frac{\|R_n\|_2}{\|L_n\|_2}. $$ But using Theorem \ref{thm:localgeom}, we find that for $n$ sufficiently large $$ \|L_n\|_2 \ge \|L_n\|_1 \ge c^\star \sum_{i=1}^{q^\star}\pi_i^n\|\theta_i^n-\theta_i^\star\|. $$ Thus we have $$ \frac{\|R_n\|_2}{\|L_n\|_2} \le \frac{d\|H_2\|_2}{2c^\star} \frac{\sum_{i=1}^{q^\star} \pi_i^n\|\theta_i^n-\theta_i^\star\|^2}{ \sum_{i=1}^{q^\star} \pi_i^n\|\theta_i^n-\theta_i^\star\|} \le \frac{d\|H_2\|_2}{2c^\star}\max_{i=1,\ldots,q^\star} \|\theta_i^n-\theta_i^\star\|\xrightarrow{n\to\infty}0. $$ We have therefore shown that $L_n/\|L_n\|_2\to d_0$ in $L^2(f^\star d\mu)$. Now define $$ \eta_i^n=\frac{\pi_i^n-\pi_i^\star}{Z_n},\quad \beta_i^n=\frac{\pi_i^n(\theta_i^n-\theta_i^\star)}{Z_n}, \qquad Z_n=\sum_{i=1}^{q^\star} \{|\pi_i^n-\pi_i^\star|+\|\pi_i^n(\theta_i^n-\theta_i^\star)\|\}. $$ As $\sum_{i=1}^{q^\star}\{|\eta_i^n|+\|\beta_i^n\|\}=1$ for all $n$, we may extract a subsequence such that $\eta_i^n\to\eta_i$, $\beta_i^n\to\beta_i$, and $\sum_{i=1}^{q^\star}\{|\eta_i|+\|\beta_i\|\}=1$. We obtain immediately $$ d_0 = \frac{L}{\|L\|_2},\qquad L= \sum_{i=1}^{q^\star}\left\{ \eta_i\,\frac{f_{\theta_i^\star}}{f^\star} +\beta_i^* \frac{D_1f_{\theta_i^\star}}{f^\star}\right\}. $$ Clearly $\sum_{i=1}^{q^\star}\eta_i=0$. Thus we have shown that any $d_0\in\mathcal{\bar D}_{q^\star}$ has the desired form. It remains to show that any function of the desired form is in fact an element of $\mathcal{\bar D}_{q^\star}$. To this end, fix $\eta_i\in\mathbb{R}$, $\beta_i\in\mathbb{R}^{d}$ with $\sum_{i=1}^{q^\star}\eta_i=0$, and define $f_t$ for $t>0$ as $$ f_t = \sum_{i=1}^{q^\star} (\pi_i^\star+t\eta_i)\, f_{\theta_i^\star+\beta_it/\pi_i^\star}. $$ Clearly $f_t\in\mathcal{M}_{q^\star}$ for all $t$ sufficiently small, and $f_t\to f^\star$ as $t\to 0$. But $$ \frac{f_t-f^\star}{t} = \sum_{i=1}^{q^\star} \pi_i^\star\,\frac{f_{\theta_i^\star+\beta_it/\pi_i^\star} -f_{\theta_i^\star}}{t} + \sum_{i=1}^{q^\star} \eta_i\,f_{\theta_i^\star+\beta_it/\pi_i^\star}. $$ Therefore clearly $$ \frac{1}{t}\frac{f_t-f^\star}{f^\star} \xrightarrow{t\to 0} \sum_{i=1}^{q^\star}\left\{ \eta_i\,\frac{f_{\theta_i^\star}}{f^\star} +\beta_i^* \frac{D_1f_{\theta_i^\star}}{f^\star}\right\} = L. $$ Using Lemma \ref{lem:chi2}, we obtain $$ \lim_{t\to 0}d_{f_t} = \lim_{t\to 0} \frac{(f_t-f^\star)/tf^\star}{ \|(f_t-f^\star)/tf^\star\|_2} = \frac{L}{\|L\|_2}. $$ Thus any function of the desired form is in $\mathcal{\bar D}_{q^\star}$, and the proof is complete. \end{proof} \begin{rem} The proof of Lemma \ref{lem:dqstar} in fact shows that $\mathcal{\bar D}_{q^\star}=\mathcal{\bar D}_{q^\star}^c$. \end{rem} We can now complete the proof of Propositions \ref{prop:mixconsist} and \ref{prop:mixconsist2}. \begin{proof}[Proof of Proposition \ref{prop:mixconsist}] We begin by proving consistency of the penalty $\mathrm{pen}(n,q)=q\,\omega(n)$. Note that by Corollary \ref{cor:localmixtures}, the assumption of Corollary \ref{cor:mainconsist} holds with $\eta(q)= 18(d+1)q+1\le 19(d+1)q$. Thus consistency of $\mathrm{pen}(n,q)=q\,\omega(n)$ follows from Corollary \ref{cor:mainconsist} using $\varpi(n) = \omega(n)/19(d+1)$. To prove inconsistency of the penalty $\mathrm{pen}(n,q)=C\,q\,\log\log n$ with $C>0$ sufficiently small, it suffices to show that $\mathcal{\bar D}_{q^\star+1}^c\backslash\mathcal{\bar D}_{q^\star}$ is nonempty. Indeed, if this is the case then we can apply Corollary \ref{cor:inconsist} with $q=q^\star+1$, where the requisite entropy assumption follows immediately from Theorem \ref{thm:mainconsist}. Fix $v\in\mathbb{R}^d$, and consider the function $f_t$ defined for $t>0$ as follows: $$ f_t = \frac{\pi_1^\star}{2}\,(f_{\theta_1^\star+vt}+ f_{\theta_1^\star-vt})+ \sum_{i=2}^{q^\star}\pi_i^\star f_{\theta_i^\star}. $$ Clearly $f_t\in\mathcal{M}_{q^\star+1}$ for all $t$ sufficiently small, $f_t\to f^\star$ as $t\to 0$, and $$ \frac{f_t-f^\star}{t^2} = \frac{\pi_1^\star}{2}\, \frac{f_{\theta_1^\star+vt}-2\,f_{\theta_1^\star} +f_{\theta_1^\star-vt}}{t^2} \xrightarrow{t\to 0} \frac{\pi_1^\star}{2}\,v^*D_2f_{\theta_1^\star}v. $$ As in the proof of Lemma \ref{lem:dqstar}, we find that $$ \lim_{t\to 0}d_{f_t} = \lim_{t\to 0} \frac{(f_t-f^\star)/t^2f^\star}{ \|(f_t-f^\star)/t^2f^\star\|_2} = \frac{v^*D_2f_{\theta_1^\star}v}{\|v^*D_2f_{\theta_1^\star}v\|_2} = d_0. $$ By construction, $d_0\in\mathcal{\bar D}_{q^\star+1}^c$. But by Theorem \ref{thm:localgeom}, the functions $f_{\theta_i^\star}$, $D_1f_{\theta_i^\star}$, and $v^*D_2f_{\theta_i^\star}v$ ($i=1,\ldots,q^\star$) are all linearly independent. Together with Lemma \ref{lem:dqstar}, this shows that $d_0\not\in\mathcal{\bar D}_{q^\star}$. Thus $d_0\in\mathcal{\bar D}_{q^\star+1}^c\backslash\mathcal{\bar D}_{q^\star}$, and the proof is complete. \end{proof} \begin{proof}[Proof of Proposition \ref{prop:mixconsist2}] By Example \ref{ex:gaussian}, the assumption of Theorem \ref{thm:mainconsist} holds with $\eta(q)= 18(d+1)q+1$ and $\log K(n) = \log C_1^\star+ C_2^\star T(n)^2$. The desired consistency results now follow immediately from Theorem \ref{thm:mainconsist}. \end{proof} \textbf{Acknowledgments.} The authors would like to thank Jean Bretagnolle for providing an enlightening counterexample which guided some of our proofs. We also thank Michel Ledoux for suggesting some helpful references.
\section{Introduction} Already the classical works of nineteenth-century geometers established a major connection between differential geometry and the theory of partial differential equations. Powerful solution-generating techniques such as the B\"acklund and Darboux transformations~\cite{R-S} have origins in the prototypical relationship between pseudospherical surfaces and solutions of the sine-Gordon equation. Methods available for solving nonlinear partial differential equations were substantially extended in the seventies to include the inverse scattering transform and its numerous developments; see, e.g.,~\cite{B-B-E-I-M,D-K-N,Ma-S,Z-M-N-P}. An important open problem is to describe the class of partial differential equations solvable by these powerful methods. Indirect detectors such as the symmetry analysis have been involved in obtaining extensive complete classifications of integrable evolution equations and systems; see~\cite{Mi-S} and references therein. The known theoretical answer given in terms of the existence of the associated one-parametric zero curvature representation $$ A_y - B_x + [A,B] = 0 $$ has been considered as a classification tool in conjunction with the gauge cohomology by one of us~\cite{spp}. These methods are not limited to evolution equations, although the necessary computations are rather complex, resource consuming, and unthinkable without substantial use of computer algebra. However, certain partial differential equations of geometric origin are particularly well suited for this classification method, namely, the Gauss--Mainardi--Codazzi equations of immersed surfaces. These equations always possess an associated linear zero curvature representation, albeit without the spectral parameter. Since their introduction by Weingarten~\cite{Wein2}, immersed surfaces in $\mathbb R^3$ that satisfy a functional relation between the principal curvatures have been of continuing interest in differential geometry, see, e.g.,~\cite{H-W,Hopf,K-S}. It is therefore not surprising that attempts have been made to identify classes of Weingarten surfaces such that the corresponding Gauss equation is integrable in the sense of soliton theory. The work of Wu~\cite{Wu} and Finkel~\cite{Fin} indicated that all integrable cases are classical, characterized by a linear relation between the Gauss and the mean curvatures (linear Weingarten surfaces~\cite[\S812]{DarIII}; see also~\cite{G-M-M} and references therein). In other words, the integrable Weingarten surfaces were conjectured to be either minimal or parallel to surfaces of constant Gaussian curvature. This conjecture was, however, disproved by the present authors in~\cite{B-M I}, henceforth referred to as Part~I. In Part I we found another integrable class, consisting of surfaces with a constant difference between the principal radii of curvature, which we called surfaces of constant astigmatism. Surprisingly enough, this extra class turned out to be classical as well, apparently first mentioned by Beltrami~\cite[Ch.~9, \S20]{Bel}, covered by Bianchi~\cite{BiaI} and Darboux~\cite{DarIII}, see also~\cite{P-Sy}, yet forgotten today. In this paper we continue the work begun in Part~I and complete the classification of integrable classes in the simplest possible case. The integrability criterion we adopt is the existence of an $\mathfrak{sl}$(2)-valued zero curvature representation depending on a nonremovable parameter. We apply the same method of formal spectral parameter, introduced in~\cite{spp} and briefly reproduced in Part~I. The underlying symbolic computations, done with the help of {\it Maple} and our own package {\it Jets}~\cite{Jets}, are omitted. To stay within the limits given by available computing resources we had to restrict the jet order (order of derivatives). The answer is given by a third order nonlinear ordinary differential equation~\eqref{ODE} to govern the functional dependence of the principal curvatures. Incorporation of the actual spectral parameter is achieved in Section~\ref{z:sect}. This can be considered a proof of integrability, opening up the possibility to obtain explicit solutions by the methods of soliton theory~\cite{B-B-E-I-M,D-K-N,Z-M-N-P}. However, we had to resign ourselves to following this road. Neither were we able to establish a B\"acklund or Darboux transformation~\cite{Ma-S,R-S}, which would allow us to construct families of exact solutions depending on an arbitrary number of parameters. We only remark that seed solutions could be conveniently found among the rotational surfaces, see~\cite[eq.~(1)]{K-S}. The governing equation~\eqref{ODE} is explored in Section~\ref{sect:ODE}. We identify two basic symmetries, scaling and translation (offsetting), and solve equation~\eqref{ODE} in terms of elliptic integrals. The generic class of integrable Weingarten surfaces we obtained depends on one essential parameter (apart from the scaling and offsetting parameters) and is believed to be new. In Sect.~\ref{sect:summary} we establish the integrable Gauss equation~\eqref{Gauss:gen} in the generic case as well as in a number of special cases when the elliptic integrals degenerate to elementary functions. All of these special cases could be located in the nineteenth century literature. Geometrically, surfaces are related by an offsetting symmetry if they are parallel to each other, i.e., if they share the same normal line congruence. Therefore, the offsetting symmetry indicates that the concept of integrability naturally extends from surfaces to their normal line congruences. Section~\ref{sect:focal} grew out of our attempt to characterize the normal congruences of the integrable Weingarten surfaces. We obtain certain relations satisfied by suitably chosen metric invariants of the pair of the focal surfaces. Naturally, we expect the corresponding focal surfaces to be integrable as well, but a detailed investigation had to be postponed to the next paper. \section{Preliminaries} \label{sect:prelim} We consider surfaces $\mathbf r(x,y)$, parameterized by the lines of curvature. This is a regular parameterization except at umbilic points. The umbilic points are isolated by the Hartman--Wintner theorem~\cite{H-W} except for spheres and planes, which are, therefore, the only surfaces excluded from consideration. The fundamental forms can be written as $$ \numbered\label{ff} `I = u^2\,`d x^2 + v^2\,`d y^2, \\ `II = \frac{u^2}\rho\,`d x^2 + \frac{v^2}\sigma\,`d y^2, $$ where $\rho,\sigma$ are the principal radii of curvature. The radii transform in a very simple way under the offsetting symmetry~\eqref{offset} of the integrability problem (unlike the principal curvatures $p = 1/\rho$, $q = 1/\sigma$ we used in Part~I). Choosing the orthonormal frame $\Psi = (\mathbf r_x/u, \mathbf r_y/v, \mathbf n)$, we consider the Gauss--Weingarten equations $$ \numbered \label{GW:so3} \Psi_x = \left(\begin{array}{ccc} \hphantom{-} 0 & -\frac{u_y}{v} & \hphantom{-} \frac u \rho \\ \hphantom{-} \frac{u_y}{v} & \hphantom{-} 0 & \hphantom{-} 0 \\ -\frac u \rho & \hphantom{-} 0 & \hphantom{-} 0 \end{array}\right) \Psi, \qquad \Psi_y = \left(\begin{array}{ccc} \hphantom{-} 0 & \hphantom{-} \frac{v_x}{u} & \hphantom{-} 0 \\ -\frac{v_x}{u} & \hphantom{-} 0 & \hphantom{-} \frac v \sigma \\ \hphantom{-} 0 & -\frac v \sigma & \hphantom{-} 0 \end{array}\right) \Psi $$ or, more explicitly, $$ \numbered\label{GW} \mathbf r_{xx} = \frac{u_x}{u} \mathbf r_x - \frac{u u_y}{v^2} \mathbf r_y + \frac{u^2}{\rho} \mathbf n, \qquad \mathbf n_{x} = -\frac 1\rho \mathbf r_x, \\ \mathbf r_{xy} = \frac{u_y}{u} \mathbf r_x + \frac{v_x}{v} \mathbf r_y, \\ \mathbf r_{yy} = -\frac{v v_x}{u^2} \mathbf r_x + \frac{v_y}{v} \mathbf r_y + \frac{v^2}{\sigma} \mathbf n, \qquad \mathbf n_{y} = -\frac 1\sigma \mathbf r_y. $$ Consequently, the Gauss--Mainardi--Codazzi equations, which are the compatibility conditions for~\eqref{GW}, read \begin{equation} \label{GMC-G} u u_{yy} + v v_{xx} - \frac{v}{u} u_x v_x - \frac{u}{v} u_y v_y + \frac{u^2 v^2}{\rho \sigma} = 0, \end{equation} and \begin{equation} \label{GMC-MC} \frac{u_y}{u} + \frac{\sigma \rho_y}{\rho (\rho - \sigma)} = 0, \quad \frac{v_x}{v} + \frac{\rho \sigma_x}{\sigma (\sigma - \rho)} = 0. \end{equation} As with Part I, we concentrate on Weingarten surfaces, which are characterized by the existence of a functional dependence between $\rho$ and $\sigma$. We often resort to a parametric representation $\rho(w),\sigma(w)$ of the dependence. Recall that parameters $x,y$ label the lines of curvature; otherwise they are arbitrary. In line with Finkel's approach~\cite{Fin}, we use this reparameterization freedom to solve the Mainardi--Codazzi subsystem~\eqref{GMC-MC}. The following proposition is a mixture of classical and new results. \begin{proposition} \label{prop:uv} Away from umbilic points, a Weingarten surface can be parameterized by the lines of curvature in such a way that $$ \numbered\label{eq:uv} u = \exp \int \frac{\rho'\sigma}{(\sigma - \rho)\rho} \,dw, \qquad v = \exp \int \frac{\rho\sigma'}{(\rho - \sigma)\sigma} \,dw. $$ The Mainardi--Codazzi subsystem~\eqref{GMC-MC} is then identically satisfied, while the remaining Gauss equation can be written in the compact form $$ \numbered\label{W:comp} R_{yy} + S_{xx} + T = 0, $$ where $R,S,T$ are appropriate functions of the unknown~$w$. Moreover, the constraint $$ \numbered\label{eq:uv1} (\frac1\rho - \frac1\sigma) u v = 1 $$ can be imposed as an additional condition, and then $T = 1/(\sigma - \rho)$. \end{proposition} \begin{proof} Writing $\rho(w),\sigma(w)$ for some function $w(x,y)$, the general solution of the Mainardi--Codazzi subsystem~\eqref{GMC-MC} is $$ u = u_0(x) \exp \int \frac{\rho'\sigma}{(\sigma - \rho)\rho} \,dw, \qquad v = v_0(y) \exp \int \frac{\rho\sigma'}{(\rho - \sigma)\sigma} \,dw. $$ Obviously from formulas~\eqref{ff}, the multipliers $u_0(x)$, $v_0(y)$ can be removed by an appropriate relabelling $\tilde x = \tilde x(x)$, $\tilde y = \tilde y(y)$ of the surface's curvature lines. With $u_0 = v_0 = 1$, we have $$ u v = \exp \int (\frac{\rho'\sigma}{(\rho - \sigma)\rho} + \frac{\rho\sigma'}{(\sigma - \rho)\sigma})\,dw = c \frac{\rho\sigma}{\sigma - \rho}, $$ where $c$ is an arbitrary constant multiplier. Setting $c = 1$ by the same relabelling argument proves the last relation. Having resolved the Mainardi--Codazzi subsystem, we are left with the Gauss equation~\eqref{GMC-G} alone. Multiplied by $1/\rho - 1/\sigma$, equation~\eqref{GMC-G} can be written in the compact form~\eqref{W:comp}, where $$ \numbered\label{RST} R = \int \frac{\rho'}{\rho^2} u^2\,dw, \qquad S = -\int \frac{\sigma'}{\sigma^2} v^2\,dw, \qquad T = u^2 v^2 \frac{\sigma - \rho}{\rho^2 \sigma^2}. $$ Substituting $1/(1/\rho - 1/\sigma)$ for $u v$ finishes the proof. \end{proof} \section{The classification result} \label{z:sect} Employing the Maple package {\it Jets}~\cite{Jets}, we completed the computer-aided cohomological classification outlined in Part~I. We have no computer-independent proof of the following result. \begin{proposition} The third-order ordinary differential equation $$ \numbered\label{ODE} \rho''' = \frac{3}{2 \rho'} \rho''{}^2 - \frac{\rho' - 1}{\rho - \sigma} \rho'' + 2 \frac{(\rho' - 1) \rho' (\rho' + 1)}{(\rho - \sigma)^2}. $$ determines a unique maximal class of Gauss--Mainardi--Codazzi equations of Weingarten surfaces whose initial $\mathfrak{sl}(2,\mathbb C)$-valued zero curvature representation $$ \numbered \label{0} A_0 = \left(\begin{array}{cc} \frac{`i u_y}{2v} & -\frac{u}{2 \rho} \\ \frac{u}{2 \rho} & -\frac{`i u_y}{2v} \end{array}\right), \qquad B_0 = \left(\begin{array}{cc} -\frac{`i v_x}{2u} & -\frac{`i v}{2 \sigma} \\ -\frac{`i v}{2 \sigma} & \hphantom{-} \frac{`i v_x}{2u} \end{array}\right) $$ admits a second order formal spectral parameter under the condition that the normal form of the zero curvature representation can depend on derivatives of $u,v,\sigma,\rho$ of no higher than the first order. \end{proposition} Here and in what follows we assume that $\rho$ is a function of $\sigma$ and the prime refers to derivatives with respect to~$\sigma$. A $k$th order formal parameter $\lambda$ means a power series in terms of $\lambda$ up to order $k$. Part~I should be consulted for the other unexplained notions. \begin{remark} \rm (1) The last proposition provides a complete classification of integrable Weingarten surfaces under the following assumptions: The one-parametric zero curvature representation takes values in the Lie algebra $\mathfrak{sl}(2)$, includes the initial zero curvature representation~\eqref{0} as a member, depends analytically on the parameter, and its normal form involves derivatives of no higher than the first order. All these limitations can be overcome, in principle~\cite{M2}, at the cost of requiring significantly more computational resources. (2) We would like to stress that the only part relying on machine computations is the completeness of the classification. All the other proofs in this paper are traditional. \end{remark} In the rest of this section we establish integrability of the class determined by equation~\eqref{ODE}. The equation itself will be solved in the next section. \begin{proposition} The nonremovable spectral parameter exists for all dependences $\rho(\sigma)$ allowed by the governing equation~\eqref{ODE}. \end{proposition} \begin{proof} Inspired by the results of the computer-aided classification, we depart from the following ansatz for the parameter-dependent zero curvature representation: $$ A = (\begin{array}{cc} a_{111} \frac{u_y} v + a_{110} \sigma_x & a_{12} u \\ a_{21} u & -a_{111} \frac{u_y} v - a_{110} \sigma_x \end{array}), \\ B = (\begin{array}{cc} b_{111} \frac{v_x} u + b_{110} \sigma_y & b_{12} v \\ b_{12} v & -b_{111} \frac{v_x} u - b_{110} \sigma_y \end{array}), $$ with $a_{111},b_{111},a_{110},b_{110},a_{12},a_{21},b_{12}$ being the unknown functions of $\sigma$. The problem is to solve the zero curvature condition $D_y A - D_x B + [A,B] = 0$ for matrix functions $A,B$ of $u,v,\sigma,\rho$ and their derivatives. However, the derivatives are not independent quantities, being subject to the Gauss--Mainardi--Codazzi equations. The proper way to deal with this situation is to introduce the manifold determined by the equation and its derivatives (a diffiety~\cite{B-V-V}). This is fairly easy if the order of derivatives is restricted as it is. Initially the derivatives are considered to be independent (jet space coordinates). Considering $\rho$ as a function of $\sigma$ and resolving the Mainardi--Codazzi equations~\eqref{GMC-MC} with respect to $u_y,v_x$, we can express $u_y,v_x$ as functions of $u,v,\sigma,\sigma_x,\sigma_y$. Similarly, the derivatives of the Mainardi--Codazzi equations~\eqref{GMC-MC} can be resolved with respect to $u_{xy},u_{yy},v_{xx},v_{xy}$, giving $u_{xy},u_{yy},v_{xx},v_{xy}$ as functions of $u,u_x,v,v_y,\sigma,\sigma_x,\sigma_y$. Consequently, the Gauss equation~\eqref{GMC-G} can be written in terms of $u,u_x,v,v_y,\sigma,\sigma_x,\sigma_y,\sigma_{xx},\sigma_{yy}$, and then resolved with respect to $\sigma_{yy}$. The explicit formulas are somewhat cumbersome, hence omitted. With $A,B$ chosen as above, the left-hand side $S := D_y A - D_x B + [A,B]$ of the zero curvature condition $S = 0$ is a matrix function of $u,u_x,v,v_y,\sigma, \sigma_x,\sigma_y,\sigma_{xx},\sigma_{xy}$. From $\partial S/\partial\sigma_{xx} = 0$ and $\partial S/\partial\sigma_{xy} = 0$ we obtain $$ b_{111} = -a_{111}, \qquad b_{110} = a_{110}. $$ From either $\partial^2 S/\partial \sigma_x^2 = 0$ or $\partial^2 S/ \partial \sigma_y^2 = 0$ we get $a_{111}' = 0$. Hence, $a_{111}$ is a constant, which we rename $\lambda$ in anticipation of its role as the spectral parameter. Now, $\partial S/\partial \sigma_x = 0$ if and only if $$ \numbered\label{b'} a_{110} = \frac{\lambda \rho}{2 \sigma (\sigma - \rho)} \, \frac{a_{12} + a_{21}}{b_{12}}, \qquad b_{12}' = \frac{\rho}{\sigma (\sigma - \rho)} [b_{12} + \lambda (a_{21} - a_{12})], $$ while $\partial S/\partial \sigma_y = 0$ can be rewritten as $$ \numbered\label{a'} a_{12}' = 2 a_{110} a_{12} + \frac{\sigma \rho'}{\rho (\rho - \sigma)} (a_{12} + 2 \lambda b_{12}), \\ a_{21}' = -2 a_{110} a_{21} + \frac{\sigma \rho'}{\rho (\rho - \sigma)} (a_{21} - 2 \lambda b_{12}), $$ Modulo these relations, vanishing of $S$ is equivalent to $$ \numbered\label{S} b_{12} = \frac{\lambda}{\rho\sigma(a_{12} - a_{21})}. $$ We claim that the governing equation~\eqref{ODE} arises as the condition that the system~\eqref{b'}, \eqref{a'}, and~\eqref{S} be compatible for arbitrary $\lambda \ne 0$. To prove this, we denote $P = a_{12} + a_{21}$, $Q = a_{12} - a_{21}$. With $a_{110}$ and $b_{12}$ taken from formulas~\eqref{b'} and \eqref{S}, respectively, equations~\eqref{a'} turn into $$ \numbered\label{P'Q'} P' = P \frac{\sigma \rho' - Q^2 \rho^3}{\rho (\rho - \sigma)}, \qquad Q' = Q \frac{\sigma \rho' - P^2 \rho^3}{\rho (\rho - \sigma)} + \frac{4 \lambda^2 \rho'}{\rho^2 (\rho - \sigma)} \, \frac1Q, $$ and the second equation in~\eqref{b'} into $$ \numbered\label{PQ} \rho^4 (Q^2 - P^2) Q^2 + \rho^2 (\rho' - 1) P^2 + 4 \lambda^2 \rho' = 0. $$ Now the question is whether equations~\eqref{P'Q'} and~\eqref{PQ} are compatible. Modulo eq.~\eqref{P'Q'}, the derivative of~\eqref{PQ} with respect to $\sigma$ is $$ \numbered\label{(PQ)'} 2 \rho^6 (P^2 - Q^2) P^2 Q^2 + 2 (1 - 3 \rho') \rho^4 P^2 Q^2 - 4 \rho' \lambda^2 P^2 \\\quad + (4 \lambda^2 + \rho^2 Q^2) [4 \rho' \rho^2 Q^2 + (\rho - \sigma) \rho'' + 2 \rho'{}^2 - 2 \rho'] = 0. $$ This is equivalent to $$ \numbered\label{Q} [(\rho - \sigma) \rho'' - 2 \rho'{}^2 + 2 (1 + 8 \lambda^2) \rho'] \rho^2 Q^2 + 4 \lambda^2 [(\rho - \sigma) \rho'' - 2 \rho'{}^2 - 2 \rho'] = 0 $$ modulo~\eqref{PQ}, since~\eqref{Q} is the remainder after division of~\eqref{(PQ)'} by~\eqref{PQ} as polynomials in~$P$. Similarly, dividing~\eqref{PQ} by~\eqref{Q} as polynomials in $Q$, we get $$ \numbered\label{P} [(\rho - \sigma) \rho'' - 2 \rho'{}^2 - 2 \rho'] [(\rho - \sigma) \rho'' - 2 \rho'{}^2 + 2 (1 + 8 \lambda^2) \rho'] \rho^2 P^2 \\\quad - 4 (1 + 4 \lambda^2) [(\sigma - \rho)^2 \rho''{}^2 - 4 \rho'{}^4 + 8 (1 + 8 \lambda^2) \rho'{}^3 - 4 \rho'{}^2] = 0. $$ Differentiating~\eqref{(PQ)'} once more and taking the result modulo~\eqref{P'Q'},~\eqref{P} and~\eqref{Q}, we get the governing equation~\eqref{ODE} immediately. Summing up, we obtain a zero curvature representation $$ A = (\begin{array}{cc} -\frac{\lambda \sigma \rho'}{\rho (\rho-\sigma)} \frac{u}{v} \sigma_y - \frac12 \frac{\rho^2}{\rho-\sigma} P Q \sigma_x & \frac12 (P + Q) u \\ \frac12 (P - Q) u & \frac{\lambda \sigma \rho'}{\rho (\rho-\sigma)} \frac{u}{v} \sigma_y + \frac12 \frac{\rho^2}{\rho - \sigma} P Q \sigma_x \end{array}), \\ B = (\begin{array}{cc} -\frac{\lambda \rho}{\sigma (\rho-\sigma)} \frac{v}{u} \sigma_x - \frac12 \frac{\rho^2}{\rho - \sigma} P Q \sigma_y & \frac{\lambda}{\sigma \rho Q} v \\ \frac{\lambda}{\sigma \rho Q} v & \frac{\lambda \rho}{\sigma (\rho-\sigma)} \frac{v}{u} \sigma_x + \frac12 \frac{\rho^2}{\rho - \sigma} P Q \sigma_y \end{array}), $$ where $P$ and $Q$ are the square roots to be determined from equations~\eqref{P} and~\eqref{Q}, respectively. Away from umbilic points (where $\rho = \sigma$), matrices $A,B$ actually exist unless $(\rho - \sigma) \rho'' - 2 \rho'{}^2 - 2 \rho' = 0$ when $P$ is undefined. This excludes exactly spheres and the linear Weingarten surfaces. The latter surfaces are, however, well known to be integrable, being parallel to surfaces of constant curvature (either Gaussian or mean), see~\cite{Wu} or \cite[\S1.5.2]{R-S}. If $\lambda = `i/2$, then we have $P = 0$ and $Q = 1/r^2$, which reproduces the parameterless zero curvature representation~\eqref{0} we started with. \end{proof} Non-removability of the parameter is ensured by the method~\cite{spp} (follows from nontriviality of the first gauge cohomology group). \section{Solution of the governing equation} \label{sect:ODE} Apart from the discrete symmetry $\rho \leftrightarrow \sigma$, the governing equation~\eqref{ODE} has two obvious continuous symmetries, which should be expected in every integrable class of surfaces: the {\it scaling symmetry} $$ \numbered\label{rescale} \rho \mapsto `e^T \rho, \qquad \sigma \mapsto `e^T \sigma $$ and the {\it translational symmetry} $$ \numbered\label{offset} \rho \mapsto \rho + T, \qquad \sigma \mapsto \sigma + T. $$ The geometric meaning of the latter symmetry is {\it offsetting\/}, also known as taking the {\it parallel surface}. In terms of position vectors, $\mathbf r$ is transformed to $\mathbf r + T \mathbf n$, where $\mathbf n$ is the unit normal vector and $T$ is the distance. With the help of these symmetries we can reduce the order of equation~\eqref{ODE} by two. This can be done by rewriting the equation in terms of the symmetry invariants. Since rescaling applies also to the offset, the translational reduction should precede the scaling reduction. For the two lowest-order translational invariants we choose $$ \numbered\label{xi eta} \xi = \rho - \sigma, \qquad \eta = \rho' $$ (recall that the prime denotes the derivative with respect to $\sigma$). \paragraph{$1$.} If $\xi' = 0$ (equivalently, $\rho' = 1$), then $\rho - \sigma = `const $, which are the surfaces of constant astigmatism we dealt with in Part~I. \paragraph{$2$.} Otherwise, more translational invariants can be computed as derivatives of $\eta$ with respect to $\xi$: $$ \numbered\label{eta_xi} \eta_\xi = \frac{\eta'}{\xi'} = \frac{\rho''}{\rho' - 1}, \quad \eta_{\xi\xi} = \frac{\rho'''}{(\rho' - 1)^2} - \frac{\rho''{}^2}{(\rho' - 1)^3}, $$ etc. In terms of these invariants, the governing equation~\eqref{ODE} reduces to the second-order equation $$ \numbered\label{ODE2} 2 \xi^2 (\eta - 1) \eta \eta_{\xi\xi} - \xi^2 (\eta - 3) \eta_\xi^2 + 2 \xi (\eta - 1) \eta \eta_\xi - 4 (\eta + 1) \eta^2 = 0. $$ As expected, this equation is scaling invariant. To reduce it with respect to scaling, we proceed as follows. Besides $\eta$, one more scaling invariant is $$ \numbered\label{zeta} \zeta = \xi (\eta - 1) \eta_\xi. $$ Although dispensable, the factor $\eta - 1$ simplifies the computations to follow. \paragraph{$2.1$.} If $\eta' = 0$, i.e., $\rho'' = 0$, then~\eqref{ODE} reduces to $\rho' = c$, where $c$ is either of $-1,0,1$. The corresponding surfaces are, respectively, the constant mean curvature surfaces (a~subclass of linear Weingarten surfaces), the tubular surfaces (surfaces swept by spheres of constant radius moving along a space curve) and once more the constant astigmatism surfaces. \paragraph{$2.2$.} Otherwise $\rho'' \ne 0$ and we have $$ \zeta_\eta = \frac{\rho'''}{\rho''} (\rho - \sigma) + \rho' - 1. $$ In terms of $\eta,\zeta$, the reduced governing equation~\eqref{ODE2} becomes the Bernoulli equation $$ \zeta_\eta = \frac32 \frac\zeta\eta + 2\frac{\eta^3 - \eta}\zeta $$ with the general solution $\zeta^2 = 4 (\eta^2 + 2 c_0 \eta + 1) \eta^2$, where $c_0$ is the integration constant. Substituting from eq.~\eqref{zeta} yields the separable first-order equation $$ \numbered\label{eta:xi} \xi \frac{d\eta}{d\xi} = \pm 2 \frac\eta{\eta - 1} \sqrt{\eta^2 + 2 c_0 \eta + 1} $$ containing the parameter $c_0$. Being written in terms of the scaling and translation invariants, this equation determines the integrable Weingarten surfaces up to rescaling and offsetting. Depending on the value of the parameter $c_0$ and on the choice of the `$\pm$' sign, we obtain the following cases. \paragraph{$2.2.1$} Let $c_0 = 1$. Equation~\eqref{eta:xi} becomes $$ \numbered\label{eta:xi 1} \xi \frac{d\eta}{d\xi} = \pm 2 \frac{(\eta + 1) \eta}{\eta - 1}. $$ \paragraph{$2.2.1.1$.} With the choice of the plus sign in~\eqref{eta:xi 1}, the general solution is $(\eta + 1)^2 = c_1 \eta \xi^2$. Substituting from eq.~\eqref{xi eta}, we obtain $$ (\rho' + 1)^2 = c_1 (\rho - \sigma)^2 \rho'. $$ If $c_1 = 0$, the general solution is $\rho + \sigma = `const$. Otherwise, we apply the transformation $$ \numbered\label{ss} \kappa = \rho + \sigma, \quad \xi = \rho - \sigma $$ to get $$ (c_1 \xi^2 - 4) (\frac{d\kappa}{d\xi})^2 = c_1 \xi^2. $$ The equation is separable with a general solution $(\kappa - c_2)^2 - \xi^2 + 4/c_1 = 0$, i.e., $$ 4 \rho\sigma - 2 c_2 (\rho + \sigma) + \frac{4 + c_1 c_2^2}{c_1} = 0. $$ In both cases, $c_1 = 0$ and $c_1 \ne 0$, solutions correspond to the linear Weingarten surfaces. \paragraph{$2.2.1.2$.} With the choice of the minus sign in~\eqref{eta:xi 1}, the general solution is $(\eta + 1)^2 \xi^2 = c_1 \eta$. Substituting from eq.~\eqref{xi eta}, we obtain $(\rho' + 1)^2 (\rho - \sigma)^2 = c_1 \rho'$. For $c_1 = 0$ we have the special linear Weingarten surfaces $\rho + \sigma = `const$ again. Otherwise, we apply the transformation \eqref{ss} to get $$ (4 \xi^2 - c_1) (\frac{d\kappa}{d\xi})^2 + c_1 = 0. $$ The solutions are $$ \kappa = \pm \frac12 \sqrt{-c_1} \ln(2 \sqrt{-c_1} \xi + \sqrt{c_1^2 - 4 c_1 \xi^2}) + c_2, $$ where $c_2$ is the integration constant. \paragraph{$2.2.1.2.1$.} For $c_1 < 0$ we can write $$ \xi = \frac{\sqrt{-c_1}}{2} \sinh(\pm\frac 2{\sqrt{-c_1}} (\kappa - c_2) - \ln(-c_1)) $$ or $$ \frac{\rho - \sigma}{C_1} = \pm \sinh(\frac{\rho + \sigma}{C_1} + C_0). $$ \paragraph{$2.2.1.2.2$.} Similarly, solutions corresponding to positive $c_1$ are $$ \numbered\label{SIN} \frac{\rho - \sigma}{C_1} = \sin(\frac{\rho + \sigma}{C_1} + C_0). $$ \paragraph{$2.2.2$} Let $c = -1$. Equation~\eqref{eta:xi} becomes $$ \numbered\label{eta:xi -1} (\eta - 1)^2 (\xi \frac{d\eta}{d\xi} - 2\eta) (\xi \frac{d\eta}{d\xi} + 2\eta) = 0. $$ Solutions corresponding to $\eta = 1$ belong to Case~1 (constant astigmatism surfaces). \paragraph{$2.2.2.1$.} The general solution of $\xi \frac{d\eta}{d\xi} = 2\eta$ is $\eta = c_1 \xi^2$. Substituting from eq.~\eqref{xi eta}, we obtain the Riccati equation $\rho' = c_1 (\rho - \sigma)^2$. \paragraph{$2.2.2.1.1$.} For $c_1 > 0$ we get $$ \numbered\label{tanh} \rho = \sigma - \frac{\tanh(\sqrt{c_1}\sigma + c_2)}{\sqrt{c_1}} \quad\text{or}\quad \rho = \sigma - \frac{\coth(\sqrt{c_1}\sigma + c_2)}{\sqrt{c_1}} $$ according to whether the integration constant is positive or negative. \paragraph{$2.2.2.1.2$.} Similarly, for $c_1 < 0$ we get $$ \numbered\label{tan} \rho = \sigma - \frac{\tan(\sqrt{-c_1}\sigma + c_2)}{\sqrt{-c_1}} \quad\text{or}\quad \rho = \sigma + \frac{\cot(\sqrt{-c_1}\sigma + c_2)}{\sqrt{-c_1}}. $$ \paragraph{$2.2.2.2$.} When solving $\xi \frac{d\eta}{d\xi} = -2\eta$, we get~\eqref{tanh} and~\eqref{tan} with $\rho,\sigma$ interchanged. \paragraph{$2.2.3$.} We are left with the generic case $c_0 \not\in \{-1,1\}$. Equation~\eqref{eta:xi} has the general solution $$ \numbered\label{generic:ode} (\eta + c_0 + \sqrt{\eta^2 + 2 c_0 \eta + 1}) (c_0 \eta + 1 + \sqrt{\eta^2 + 2 c_0 \eta + 1}) = c_1 \xi^{\pm 2} \eta. $$ If $c_1 = 0$, then $\eta = 0$ in view of $c_0 \not\in \{-1,1\}$, which yields the tubular surfaces $\rho = `const$. Let us, therefore, assume that $c_1 \ne 0$. Upon substituting from~\eqref{xi eta}, equation~\eqref{generic:ode} becomes a first-order ODE, separable in terms of variables~\eqref{ss} and having the elliptic integral $$ \kappa = \int^\xi \frac{-c_1 t^{\pm2} + c_0^2 - 1} {\sqrt{c_1^2 t^{\pm4} - 2 (c_0 + 1) (c_0 + 3) c_1 t^{\pm2} + (c_0^2 - 1)^2}} \,`d t $$ as the general solution. The two cases the `$\pm$' symbol refers to can be converted one into another by the substitution $c_1 \to (c_0^2 - 1)^2/c_1$. Therefore, we can safely choose the sign to be `$+$', which we do in the sequel. Moreover, if $\kappa$ is a solution, then so is $-\kappa$ (as a combination of the $\rho \leftrightarrow \sigma$ switch and a scaling by factor of $-1$). This is why we often ignore the sign of~$\kappa$ in what follows. Substituting $t \to s/m$, $m = \sqrt{|c_1/(1 - c_0^2)|}$, we simplify the integral above to $$ \numbered\label{Ipm} \kappa = \frac 1m I_\pm(m\xi,c), \qquad I_\pm(\xi,c) = \int^{\xi} \frac{1 \pm s^2} {\sqrt{1 + 2 c s^2 + s^4}} \,`d s, $$ where `$\pm$' refers to the signum of $c_1/(1 - c_0^2)$; in particular, is unrelated to the `$\pm$' sign in~\eqref{generic:ode}. The real parameter $c$ is related to $c_0$ by $c = \pm\frac{c_0 + 3}{c_0 - 1}$. Formula~\eqref{Ipm} describes possible dependences $\rho(\sigma)$ via the substitution $\kappa = \rho + \sigma$, $\xi = \rho - \sigma$. Three independent parameters are involved: $m,c$ and the integration constant (the lower limit of the integral). Obviously, $m$ plays the role of the scaling parameter. The integration constant can be easily identified with the offsetting parameter $T$ from~\eqref{offset}. Each dependence between $\kappa$ and $\xi$ has a unique representative modulo scaling and offsetting, obtainable by fixing the lower limit of the integral $I_\pm(\xi,c)$ in~\eqref{Ipm}. This is straightforward when $c > -1$; we simply redefine $I_\pm(\xi,c)$ to be $$ \numbered\label{Ipm_0} I_\pm(\xi,c) = \int^{\xi}_0 \frac{1 \pm s^2} {\sqrt{1 + 2 c s^2 + s^4}} \,`d s. $$ If, however, $c < -1$, then the integrand in~\eqref{Ipm} is real in three separate intervals $(-\infty,-\sqrt{\gamma_+})$, $(-\sqrt{\gamma_-},\sqrt{\gamma_-})$, and $(\sqrt{\gamma_+},\infty)$, where $$ \numbered\label{gamma(c)} \gamma_\pm = -c \pm \sqrt{c^2 - 1} > 0. $$ We choose the representatives $-\tilde I_\pm(-\xi,c)$, $I_\pm(\xi,c)$, and $\tilde I_\pm(\xi,c)$, respectively, where $I_\pm(\xi,c)$ is given by~\eqref{Ipm_0} in the interval $-\gamma_- \le \xi \le \gamma_-$, while $$ \numbered\label{tilde Ipm} \tilde I_\pm(\xi,c) = \int^{\xi}_{\gamma_+} \frac{1 \pm s^2} {\sqrt{1 + 2 c s^2 + s^4}} \,`d s, \quad \gamma_+ \le \xi. $$ \section{Summary of the solutions} \label{sect:summary} As demonstrated in the preceding section, each integrable class is determined by certain relation between the radii of curvature, which can be subject to rescaling $\rho \to c_1 \rho$, $\sigma \to c_1 \sigma$, offsetting $\rho \to \rho + c_0$, $\sigma \to \sigma + c_0$ and the twist $\rho \leftrightarrow \sigma$. With the help of Proposition~\ref{prop:uv}, we can find the corresponding integrable Gauss equation. To start with, we investigate the generic class determined by formula~\eqref{Ipm}; we fix the scaling for simplicity. \begin{proposition} Assuming $$ \numbered\label{gen} \rho + \sigma = I_{\pm}(\rho - \sigma,c), \quad I_{\pm}(\xi,c) = \int^\xi \frac{1 \pm s^2}{\sqrt{1 + 2 c s^2 + s^4}}\,`d s, $$ the Gauss equation~\eqref{GMC-G} for $\xi = \rho - \sigma$ reads $$ \numbered\label{Gauss:gen} R' \xi_{yy} + R'' \xi_y^2 + S' \xi_{xx} + S'' \xi_x^2 + T = 0, $$ where $$ R' = \frac{1 + c \xi^2 + \Delta(\xi,c)}{\xi^2 \Delta(\xi,c)}, \quad S' = \frac{c \mp 1}2\,\frac{\xi^2}{(1 + c \xi^2 + \Delta(\xi,c)) \Delta(\xi,c)}, \\ \Delta(\xi,c) = \sqrt{1 + 2 c \xi^2 + \xi^4}, \quad T = -\frac 1\xi. $$ The metric coefficients $u,v$ in~\eqref{ff} are $$ u = \frac{\xi + I_{\pm}(\xi,c)}{2 \xi} \sqrt{1 \mp \xi^2 + \Delta(\xi,c)}, \quad v = \frac{\xi - I_{\pm}(\xi,c)}{2 \xi} \sqrt{\frac{1 \mp \xi^2 - \Delta(\xi,c)}{2 c \pm 2}}. $$ \end{proposition} \begin{proof} We parameterize $\rho$ and $\sigma$ by $\xi$, i.e., we resolve~\eqref{gen} as $$ \rho = \frac{I_{\pm}(\xi,c) + \xi}2, \quad \sigma = \frac{I_{\pm}(\xi,c) - \xi}2. $$ The general form of the Gauss equation, along with the last term $T = 1/(\sigma - \rho) = -1/\xi$, follow from Proposition~\ref{prop:uv}. To find $R',S'$, we compute $$ (\ln R')' = \frac{R''}{R'} = \frac{(\rho - \sigma)\rho'' - 2\rho^{\prime2}}{(\rho - \sigma)\rho'} = -\frac2\xi \, \frac{c \xi^2 + \xi^4 + \sqrt{1 + 2 c \xi^2 + \xi^4}}{1 + 2 c \xi^2 + \xi^4}, \\ (\ln S')' = \frac{S''}{S'} = \frac{(\rho - \sigma)\sigma'' + 2\sigma^{\prime2}}{(\rho - \sigma)\sigma'} = -\frac2\xi \, \frac{c \xi^2 + \xi^4 - \sqrt{1 + 2 c \xi^2 + \xi^4}}{1 + 2 c \xi^2 + \xi^4} $$ from~\eqref{RST} under the constraint~\eqref{eq:uv1}. These equations need to be integrated once, which is easy; the integration constants have been chosen to match equations~\eqref{eq:uv1} and~\eqref{RST}. Finally, from~\eqref{RST} one easily computes the coefficients $u,v$ as $u = \sqrt{{R' \rho^2}/{\rho'}}$, $v = \sqrt{-{S' \sigma^2}/{\sigma'}}$. \end{proof} Apart from the generic class we also obtained a number of special solutions, listed in Table~\ref{T1} (omitting the tubular surfaces). Rows 5b and 6b differ only by translation (offsetting) and can be identified one with another. The first column contains a determining relation (up to a scaling), while the second harbours the corresponding integrable equation in the compact form~\eqref{W:comp}. Table~\ref{T2} gives the principal radii of curvature $\rho,\sigma$, metric coefficients $u,v$, and the variable $z$ (see Table~\ref{T1}) in terms of a suitably chosen parameterizing variable $w$. \begin{table} $$ \begin{array}{lll} & \text{relation} & \text{integrable equation} \\ \hline 1. & \rho + \sigma = 0 & z_{xx} + z_{yy} + `e^z = 0 \\ \text{2a}. & \rho\sigma = 1 & z_{xx} + z_{yy} - \sinh z = 0 \\ \text{2b}. & \rho\sigma = -1 & z_{xx} - z_{yy} + \sin z = 0 \\ \text{3a}. & \rho - \sigma = \sinh(\rho + \sigma) & (\tanh z - z)_{xx} + (\coth z - z)_{yy} + `csch 2 z = 0 \\ \text{3b}. & \rho - \sigma = \sin(\rho + \sigma) & (\tan z - z)_{xx} + (\cot z + z)_{yy} + \csc 2 z = 0 \\ 4. & \rho - \sigma = 1 & z_{xx} + (1/z)_{yy} + 2 = 0 \\ \text{5a}. & \rho - \sigma = \tanh \rho & \frac14(\sinh z - z)_{xx} + (\coth \frac12 z)_{yy} + \coth \frac12 z = 0 \\ \text{5b}. & \rho - \sigma = \tan \rho & \frac14(\sin z - z)_{xx} + (\cot \frac12 z)_{yy} + \cot \frac12 z = 0 \\ \text{6a}. & \rho - \sigma = \coth \rho & \frac14(\sinh z + z)_{xx} - (\tanh \frac12 z)_{yy} + \tanh \frac12 z = 0 \\ \text{6b}. & \rho - \sigma = -\cot \rho & \frac14(\sin z + z)_{xx} + (\tan \frac12 z)_{yy} + \tan \frac12 z = 0 \end{array} $$ \caption{Special integrable cases and the associated integrable Gauss equations} \label{T1} \end{table} \begin{table} $$ \begin{array}{lccccc} & \rho & \sigma & u & v & z \\ \hline 1. & w & \llap{$-$}w & \sqrt{w/2} & \sqrt{w/2} & \llap{$-$}\ln w \\ \text{2a}. & w & \frac 1w & \frac{w}{\sqrt{w^2 - 1}} & \frac {-1}{\sqrt{w^2 - 1}} & 2 `arctanh w \\ \text{2b}. & w & \llap{$-$}\frac 1w & \frac{w}{\sqrt{w^2 + 1}} & \frac 1{\sqrt{w^2 + 1}} & 2 `arctan w \\ \text{3a}. & \frac{w + \sinh w}2 & \frac{w - \sinh w}2 & \frac{w + \sinh w}{2 \sqrt{\cosh w - 1}} & \frac{w - \sinh w}{2 \sqrt{\cosh w + 1}} & \frac 12 w \\ \text{3b}. & \frac{w + \sin w}2 & \frac{w - \sin w}2 & \frac{w + \sin w}{2 \sqrt{1 - \cos w}} & \frac{w - \sin w}{2 \sqrt{1 + \cos w}} & \frac 12 w \\ 4. & w & w - 1 & \frac{w}{`e^{w}} & (1 - w) `e^{w} & `e^{2w} \\ \text{5a}. & w & w - \tanh w & \frac{w}{\sinh w} & \sinh w - w \cosh w & 2 w \\ \text{5b}. & w & w - \tan w & \frac{w}{\sin w} & \sin w - w \cos w & 2 w \\ \text{6a}. & w & w - \coth w & \frac{w}{\cosh w} & \cosh w - w \sinh w & 2 w \\ \text{6b}. & w & w + \cot w & \frac{w}{\cos w} & \cos w + w \sin w & 2 w \end{array} $$ \caption{Special integrable cases. The radii of curvature $\rho, \sigma$, the metric coefficients $u, v$, and the unknown $z$ of the integrable Gauss equation in terms of a variable $w$.} \label{T2} \end{table} Neither of the special cases is new to differential geometry. Row 1 reflects that, in terms of the curvature line coordinates, minimal surfaces correspond to solutions of the Liouville equation~\cite[\S351]{BiaII}. Similarly, row 2a reproduces the relation between surfaces of negative constant Gaussian curvature and solutions of the elliptic sinh-Gordon equation. Row 2b does the same for the hyperbolic sine-Gordon equation and surfaces of positive constant Gaussian curvature (or constant mean curvature, by the theorem of Bonnet on parallel surfaces). Nowadays, surfaces of constant mean or Gaussian curvature are undoubtedly the best understood classes of surfaces integrable in the sense of soliton theory (see, e.g.,~\cite{Bob1,Bob4,D-N,Hel,Me-S,P-St} and references therein). It may come as a surprise that the other cases are classical as well. Introduced by Weingarten~\cite[\S4]{Wein2} (`eine neue Fl\"achenklasse'), surfaces satisfying the relation $\rho - \sigma = \sin(\rho + \sigma)$ (row 4b) are covered in Darboux~\cite[\S\S745, 746, 766, 769, 770]{DarIII} (`une classe nouvelle de surfaces d\'ecouverte par M.~Weingarten') and Bianchi~\cite[\S135]{BiaI}, \cite[\S245]{BiaII}. Darboux~\cite[\S746]{DarIII} gave a general solution of an equation equivalent to our $(\tan z - z)_{xx} + (\cot z + z)_{yy} + \csc 2 z = 0$. He also provided a remarkable geometric construction in~\cite[\S770]{DarIII}, further developed by Bianchi~\cite[\S245]{BiaII}. In a nutshell: the middle evolutes are translation surfaces generated by curves of opposite constant nonzero torsion; conversely the Weingarten surfaces are orthogonal to the osculation planes of the generating curves. Bianchi's research extends to the complementary relation $\rho - \sigma = \sinh(\rho + \sigma)$ (row 3a) as well~\cite[\S246]{BiaII}. The remaining rows (from 4 to 6b) correspond to involutes of surfaces of constant Gaussian curvature studied by Beltrami~\cite[Ch.~9, \S20]{Bel}. Row~4 (surfaces of constant astigmatism) has been addressed in Part~I; we have nothing to add except the Beltrami's work as the earliest reference we know of. Table~\ref{Tlim} demonstrates how the cases expressible in terms of elementary functions arise as limits of the generic integral~\eqref{Ipm} for $c$ approaching $\pm 1$ or $\pm\infty$ along a suitable curve in the $(c,m)$ space. The tubular surfaces $\sigma = `const$, which are omitted, correspond to $\kappa = I_+(\xi,1) = \xi + `const$. \begin{table} $$ \begin{array}{lll} & \text{relation} & \text{limit} \\ \hline 1. & \kappa = 0 & I_\pm (\xi, \infty) \\ \text{2a}. & \kappa^2 = \xi^2 + 4 & \lim_{m = \infty} I_\pm (m\xi, 2 m^2)/m \\ \text{2b}. & \kappa^2 = \xi^2 - 4 & \lim_{m = \infty} I_\pm (m\xi, -2 m^2)/m \\ \text{3a}. & \kappa = `arcsinh \xi & \lim_{m = 0} I_\pm (m\xi, 1/{2 m^2})/m \\ \text{3b}. & \kappa = `arcsin \xi & \lim_{m = 0} I_\pm (m\xi, -1/{2 m^2})/m \\ 4. & \xi = 1 & \lim_{m = \infty} \tilde I_\pm (m\xi, -m^2/2)/m \\ \text{5a}. & \kappa = -\xi + 2`arctanh \xi & I_+(\xi,-1),\ |\xi| \lt 1 \\ \text{5b}. & \kappa = -\xi + 2`arctan \xi & I_-(\xi,1) \\ \text{6a}. & \kappa = -\xi + 2`arccoth \xi & I_+(\xi,-1),\ |\xi| \gt 1 \\ \text{6b}. & \kappa = -\xi - 2`arccot \xi & I_-(\xi,1) \end{array} $$ \caption{Special integrable cases as limits of $I_{\pm}(\xi,c)$} \label{Tlim} \end{table} \section{Curvature diagrams} To exemplify the wealth of classes of integrable surfaces, we plot the representative solutions of the governing equation~\eqref{ODE} in Figures~1 and~2. We call them curvature diagrams, even though the radii of curvature $\rho,\sigma$, rather than the curvatures $1/\rho,1/\sigma$, are plotted, contrary to the customary practice~\cite[Ch.~5]{H-C}. The benefit is that diagrams can be not only scaled arbitrarily, but also freely translated along the dashed line $\rho = \sigma$; the translation corresponds to offsetting. For clarity, we adjusted the offsetting so that the diagrams are symmetric about the origin, i.e., $\rho(\sigma) = -\rho(-\sigma)$. The diagrams contain plots of functions $\mathcal I_{`A}(\xi,k)$, $\mathcal{I}_{`B}(\xi,k)$, $\mathcal{I}_{`C\pm}(\xi,k)$, and $k\tilde{\mathcal I}_{`A}(\xi/k,k)$. All special cases are explicitly included as limits, except the surfaces of constant positive curvature (row~2a). These could be obtained as the limit of $k \mathcal{I}_{`B}(\xi/k, k)$ as $k$ approaches zero. \begin{figure} \begin{center} \includegraphics[width=2\unitlength]{figure1.pdf} \caption{Curvature diagrams $\kappa = \mathcal{I}_{`B}(\xi, k)$ (the left-hand legend) and $\kappa = \mathcal{I}_{`A}(\xi,k)$, $|\xi| < 1$ (the right-hand legend), where $\kappa = \rho+\sigma$, $\xi = \rho - \sigma$. More can be obtained by rescaling and translating along the dashed line $\rho = \sigma$, the axis $\kappa$. Here $\mathcal{I}_{`A}(\xi,-1) = -\xi + 2 `arctan \xi$ (row~5b), $\mathcal{I}_{`A}(\xi,0) = `arcsin \xi$ (row~3b), $\mathcal{I}_{`A}(\xi,1) = \xi$; $\mathcal{I}_{`B}(\xi,-1) = -\xi + 2 `arctanh \xi$ (row~5a), $\mathcal{I}_{`B}(\xi,0) = `arcsinh \xi$ (row~3a), $\mathcal{I}_{`B}(\xi,1) = \xi$. Graphs of $\kappa = \mathcal{I}_{`A}(\xi,k)$ end on the solid lines $|\xi| = 1$.} \end{center} \end{figure} \begin{figure} \begin{center} \begin{picture}(1,1)(0,0) \put(0.5,0.5){\makebox(0,0){\includegraphics[width=\unitlength]{figure2a.pdf}}} \put(0,0){\makebox(0,0)[lb]{\scriptsize(a)}} \end{picture} \quad \begin{picture}(1,1)(0,0) \put(0.5,0.5){\makebox(0,0){\includegraphics[width=\unitlength]{figure2b.pdf}}} \put(0,0){\makebox(0,0)[lb]{\scriptsize(b)}} \end{picture} \end{center} \caption{Curvature diagrams (a) $\kappa = k \tilde{\mathcal{I}}_{`A}(\xi/k, k)$, $|\xi|> 1/|k|$; (b) $\kappa = \mathcal{I}_{C+}(\xi,k)$ (the top left-hand legend) and $\kappa = -\mathcal{I}_{C-}(\xi,k)$ (the bottom right-hand legend), where $\kappa = \rho+\sigma$, $\xi = \rho - \sigma$. More can be obtained by rescaling and translating along the dashed line $\rho = \sigma$, the axis $\kappa$. \\ In (a), the line $k = 1$ corresponds to tubular surfaces, $k = 0$ to surfaces of negative constant curvature (row~2b), and $k = -1$ to the constant astigmatism surfaces (row~4). In (b), $\mathcal{I}_{C+}(\xi,0) = -\xi + 2 `arctan \xi$ (row~5b) $\mathcal{I}_{C-}(\xi,1) = \xi - 2`arctan((\xi - 1)/(\xi + 1))$ (row~5a after reparameterization). } \end{figure} The plots have been calculated using the Legendre normal form~\cite{Erd,P-So} of the elliptic integrals~\eqref{Ipm_0} and~\eqref{tilde Ipm}, which could be of independent interest. As is well known, the Legendre normal form depends on the configuration of roots of the quartic polynomial $\Pi = s^4 + 2 c s^2 + 1$. A) If $c < -1$, then $\Pi = (s^2 - \gamma_+)(s^2 - \gamma_-)$ has four real roots $\sqrt{\gamma_\pm}$ and $-\sqrt{\gamma_\pm}$ given by formula~\eqref{gamma(c)}. By using the substitution $s = \sqrt{k} r$, where $k = \gamma_-$, we easily obtain the Legendre normal form $$ \frac1{\sqrt{k}} I_\pm(\xi \sqrt{k},-\frac{k^2 + 1}{2 k}) = \int_0^{\xi} \frac{1 \pm k r^2}{\sqrt{(1 - r^2)(1 - k^2 r^2)}}\,`d r, \quad 0 < k < 1. $$ On the right-hand side, we can remove the $\pm$ sign from the numerator by allowing $k$ to range between $-1$ and~$1$. For $-1 \le \xi \le 1$, $-1 < k < 1$, we have a unified representative given by $\kappa = \mathcal I_{`A}(\xi,k)$, where $$ \mathcal I_{`A}(\xi,k) = \int_0^{\xi} \frac{1 - k r^2}{\sqrt{(1 - r^2)(1 - k^2 r^2)}}\,`d r = \frac{1}{k} E(\xi; k) + \frac{k - 1}{k} F(\xi; k) $$ in terms of the Legendre elliptic integrals $E,F$. For real $\xi$ such that $|\xi| > 1$, the function $\mathcal I_{`A}(\xi,k)$ is complex valued. Yet we obtain a real function for $1/|k| \le \xi$ by choosing the lower limit of the integral to be $1/k$, $-1 < k < 1$. Thus, $$ \tilde{\mathcal I}_{`A}(\xi,k) = \begin{cases} \int_{1/|k|}^{\xi} \frac{1 - k r^2}{\sqrt{(1 - r^2)(1 - k^2 r^2)}}\,`d r = \mathcal I_{`A}(\xi,k) - \mathcal I_{`A}(\frac1{|k|},k), & \xi > \frac1{|k|}, \\ -\tilde{\mathcal I}_{`A}(-\xi,k), & \xi < -\frac1{|k|}. \end{cases} $$ B) Similarly, when $c > 1$, then $\gamma_\pm \lt 0$, the roots $\sqrt{\gamma_\pm}$, $-\sqrt{\gamma_\pm}$ of $\Pi$ are purely imaginary, and $$ \frac1{\sqrt{k}} I_\pm(\xi \sqrt{k}, \frac{k^2 + 1}{2 k}) = \int_0^{\xi} \frac{1 \pm k r^2}{\sqrt{(1 + r^2)(1 + k^2 r^2)}}\,`d r, \quad 0 < k < 1. $$ The two representatives can be unified into $\kappa = \mathcal I_{`B}(\xi,k)$, where $$ \mathcal I_{`B}(\xi,k) = \int_0^{\xi} \frac{1 - k r^2}{\sqrt{(1 + r^2)(1 + k^2 r^2)}}\,`d r = \frac{1}{k`i} E(\xi`i; k) + \frac{k - 1}{k`i} F(\xi`i; k) $$ for $-1 < k < 1$. C) When $-1 < c < 1$ (four distinct complex roots), we substituted $$ s = \frac{1 + \sqrt{k} r}{1 - \sqrt{k} r}, \quad 0 < k < 1, $$ to obtain two more representatives $\kappa = \mathcal I_{`C+}$ and $\kappa = \mathcal I_{`C-}$, where $$ \mathcal I_{`C\pm} = \begin{cases} J_{`C\pm}(\xi,k) - J_{`C\pm}(0,k), & \xi \ge 0, \\ -\mathcal I_{`C\pm}(-\xi,k), & \xi \lt 0, \end{cases} \\ J_{`C\pm}(\xi,k) = \frac{\sqrt{1 + 2 c \xi^2 + \xi^4}}{1 + \xi} + \frac{2}{(k + 1)`i} E(\frac{\xi - 1}{\xi + 1}\frac{`i}{\sqrt k}, k) + \frac{\epsilon_\pm}{`i} F(\frac{\xi - 1}{\xi + 1}\frac{`i}{\sqrt k}, k), \\ \epsilon_\pm = \frac{(1 \pm 1)k - 3 \pm 1}{2} = \begin{cases} k - 1, \\ -2, \end{cases} \\ c = -\frac{k^2 - 6 k + 1}{(k+1)^2}. $$ \section{Normal congruences and their focal surfaces} \label{sect:focal} The fact that the governing equation~\eqref{ODE} has the offsetting symmetry~\eqref{offset} is not a pure coincidence. Being invertible, the offsetting transformation $\mathbf r \mapsto \mathbf r + T \mathbf n$ preserves integrability in every reasonable sense of the word. Surfaces related by the offsetting transformation are said to be parallel and either all are integrable or none is. However, parallel surfaces can be alternatively described as normal surfaces to the same line congruence. Consequently, integrability is a property of this congruence and, therefore, must have an expression in terms of congruence invariants. Normal congruences of Weingarten surfaces, also known as $W$-congruences, are rather special with regard to properties of their focal surfaces. It is therefore natural to look for characterization of the former in terms of the latter. Naturally, we expect the focal surfaces of integrable $W$-congruences to be integrable as well. Recall that a generic surface has two focal surfaces (often considered as two sheets of a single surface), $$ \mathbf r^{(1)} = \mathbf r + \sigma \mathbf n, \qquad \mathbf r^{(2)} = \mathbf r + \rho \mathbf n. $$ each of which is formed by the evolutes of one family of the curvature lines. Focal surfaces can degenerate into a line or even a point. In the case of a Weingarten surface $\mathbf r$ with fundamental forms~\eqref{ff}, one of the focal surfaces degenerates into a line if $\sigma_y = \sigma' w_y = 0$ or $\rho_x = \rho' w_x = 0$; both degenerate into a point if the surface is a sphere (already excluded from consideration); otherwise they are regular surfaces. Therefore, we assume $\rho' \sigma' \ne 0$ in what follows. To compute the respective first and second fundamental forms $`I^{(i)}$ and $`II^{(i)}$, $i = 1,2$, we proceed as follows. In view of the Gauss--Mainardi--Codazzi equations~\eqref{GMC-G} and~\eqref{GMC-MC}, the Gauss--Weingarten~\eqref{GW} equations can be written as $$ \numbered \label{GW*} \mathbf r_{xx} = \frac{u_x}{u} \mathbf r_x + \frac{\sigma \rho' u^2 w_y}{\rho (\rho - \sigma) v^2} \mathbf r_y + \frac{u^2}{\rho} \mathbf n, \qquad \mathbf n_x = -\frac1\rho \mathbf r_x, \\ \mathbf r_{xy} = \frac{\sigma \rho' w_y}{\rho(\sigma - \rho)} \mathbf r_x + \frac{\rho \sigma' w_x}{\sigma(\rho - \sigma)} \mathbf r_y, \\ \mathbf r_{yy} = \frac{\rho \sigma' v^2 w_x}{\sigma (\sigma - \rho) u^2} \mathbf r_x + \frac{v_y}{v} \mathbf r_y + \frac{v^2}{\sigma} \mathbf n, \qquad \mathbf n_y = -\frac1\sigma \mathbf r_y. $$ One easily finds $$ \mathbf r^{(1)}_x = \frac{\rho - \sigma}\rho \mathbf r_x + \sigma' w_x \mathbf n, \quad \mathbf r^{(1)}_y = \sigma' w_y \mathbf n, \quad \mathbf n^{(1)} = \frac{\mathbf r_y}{v}, \\ \mathbf r^{(2)}_x = \rho' w_x \mathbf n, \quad \mathbf r^{(2)}_y = \frac{\sigma - \rho}\sigma \mathbf r_y + \rho' w_y \mathbf n, \quad \mathbf n^{(2)} = \frac{\mathbf r_x}{u}. $$ Using the equations~\eqref{GW*} and~\eqref{ff}, we get $$ \numbered\label{evol:ff1} `I^{(1)} = \frac{(\rho - \sigma)^2 u^2}{\rho^2}`d x^2 +\, `d\sigma^2, \qquad `I^{(2)} = d\rho^2 + \frac{(\rho - \sigma)^2 v^2}{\sigma^2}`d y^2, $$ where $`d\rho = \rho'\,`d w = \rho'(w_x\,`d x + w_y\,`d y)$, $`d\sigma = \sigma'\,`d w = \sigma'(w_x\,`d x + w_y\,`d y)$. With $u,v$ determined from Proposition~\ref{prop:uv}, we can write $$ `I^{(1)} = (f^{(1)}(\sigma)\,`d x)^2 +\, `d\sigma^2, \qquad `I^{(2)} = (f^{(2)}(\rho)\,`d y)^2 +\, `d\rho^2. $$ Hence, all focal surfaces $\mathbf r^{(i)}$ corresponding to a given dependence $\rho(\sigma)$ are isometric. Moreover, the first fundamental forms~\eqref{evol:ff1} are typical of surfaces of revolution. These are among the classical results by Weingarten~\cite{Wein2}. Omitting details, we further compute the second fundamental forms $$ \numbered\label{evol:ff2} `II^{(1)} = \frac{\sigma w_y}{v} (\frac{\rho' u^2}{\rho^2}`d x^2 - \frac{\sigma' v^2}{\sigma^2}`d y^2), \quad `II^{(2)} = -\frac{\rho w_x}{u} (\frac{\rho' u^2}{\rho^2}\,`d x^2 - \frac{\sigma' v^2}{\sigma^2}`d y^2) $$ and note that they are conformally related, which is another way to express Ribaucour's classical result~\cite{Ri} that asymptotic coordinates on $\mathbf r^{(1)}$ and $\mathbf r^{(2)}$ correspond. The Gaussian curvatures are $$ \numbered\label{evol:GC} K^{(1)} = \frac{\det `II^{(1)}}{\det `I^{(1)}} = -\frac{\rho'}{(\rho - \sigma)^2 \sigma'}, \qquad K^{(2)} = \frac{\det `II^{(2)}}{\det `I^{(2)}} = -\frac{\sigma'}{(\rho - \sigma)^2 \rho'}. $$ Consequently, the focal surfaces have one and the same sign of the Gaussian curvature, which we denote as $\epsilon$. We have $\epsilon = -1$ (both focal surfaces are hyperbolic) if and only if $`d\rho/`d\sigma = \rho'/\sigma' > 0$ (if $\rho$ increases as $\sigma$ increases), and $+1$ if $`d\rho/`d\sigma < 0$. The relation $$ \numbered\label{Hal} K^{(1)} K^{(2)} = \frac 1{(\rho - \sigma)^4} $$ away of umbilic points is known as the Halphen theorem (see~\cite[\S129]{BiaI}). As we have already explained, to every particular relation $\rho(\sigma)$ of curvatures there corresponds an isometry class of focal surfaces, which contains a unique rotational representative (which is the way the classes have been characterized in the classical literature). However, we believe that a description in terms of metric invariants is more appropriate. It is convenient to choose $$ \kappa^{(i)} = \frac1{\sqrt{\epsilon K^{(i)}}}, $$ where $\epsilon K^{(i)} = |K^{(i)}|$ is the absolute value of the Gaussian curvature of the $i$th focal surface. Further, let $\gamma^{(i)}$ be defined by $$ \numbered\label{gamma12} \gamma^{(1)} = \frac{(\rho - \sigma) (\rho'' \sigma' - \sigma'' \rho') - 2 \rho' \sigma' (\rho' - \sigma')}{2\,(-\epsilon \rho' \sigma')^{3/2}}, \\ \gamma^{(2)} = \frac{(\rho - \sigma) (\rho'' \sigma' - \sigma'' \rho') + 2 \rho' \sigma' (\rho' - \sigma')}{2\,(-\epsilon \rho' \sigma')^{3/2}}. $$ One can directly check that $|\gamma^{(i)}|$ equals the norm of the gradient of $\kappa^{(i)}$ with respect to~$`I^{(i)}$, $$ |\gamma^{(i)}| = \left\|`grad^{(i)}\kappa^{(i)}\right\|^{(i)} = \sqrt{I^{(i)} (`grad^{(i)}\kappa^{(i)}, `grad^{(i)}\kappa^{(i)})}. $$ Hence, $\gamma^{(i)}$ is a metric invariant of the respective focal surface. It is sometimes more convenient to use invariants $$ \numbered\label{G12} G^{(1)} = \frac{[(\rho - \sigma) (\rho'' \sigma' - \sigma'' \rho') - 2 \rho' \sigma' (\rho' - \sigma')]^2}{16\,(\rho' \sigma')^3}, \\ G^{(2)} = \frac{[(\rho - \sigma) (\rho'' \sigma' - \sigma'' \rho') + 2 \rho' \sigma' (\rho' - \sigma')]^2}{16\,(\rho' \sigma')^3}, $$ satisfying $$ \gamma^{(i)2} = -4\epsilon G^{(i)}, \quad -16\,G^{(i)} K^{(i)3} = I^{(i)} (`grad^{(i)} K^{(i)}, `grad^{(i)} K^{(i)}). $$ Clearly, both $\kappa^{(i)}$ and $G^{(i)}$ are functions of $w$. Consequently, $G^{(i)}$ can be considered as a function of $\kappa^{(i)}$ unless $\kappa^{(i)}$ is a constant. Our nearest aim is to establish the dependence between $\kappa^{(i)}$ and $G^{(i)}$ in terms of the dependence between $\rho$ and $\sigma$. \begin{proposition} \label{prop:focal} Let the principal radii of curvature $\rho,\sigma$ of an integrable surface satisfy the generic relation~\eqref{Ipm}. Then the metric invariants $G^{(i)}$ and $\kappa^{(i)}$ satisfy the relations $$ \numbered\label{Gkappa} G^{(1)} = (-1 \pm \sqrt{\frac 2{c \mp 1}} \frac{\kappa^{(1)}}m) (1 + \sqrt{\frac 2{c \mp 1}} \frac m{\kappa^{(1)}}), \\ G^{(2)} = (1 \pm \sqrt{\frac 2{c \mp 1}} \frac{\kappa^{(2)}}m) (-1 + \sqrt{\frac 2{c \mp 1}} \frac1 m{\kappa^{(2)}}). $$ Furthermore, $$ G^{(1)} G^{(2)} = (\frac{c \pm 1}{c \mp 1})^2 $$ is constant (hence, so is the product $\gamma^{(1)} \gamma^{(2)}$). Table~\ref{T4} lists the product $G^{(1)} G^{(2)}$ and the algebraic relations between $G^{(i)}$ and $\kappa^{(i)}$ in the special cases. \end{proposition} \begin{proof} For simplicity, we start assuming a fixed scaling, i.e., we depart from formula~\eqref{gen}. We routinely compute $$ K^{(1)} = \frac{(1 \pm w^2 + \sqrt{1 + 2c w^2 + w^4})^2} {2 (c \mp 1) w^4}, \quad K^{(2)} = \frac{(1 \pm w^2 - \sqrt{1 + 2c w^2 + w^4})^2} {2 (c \mp 1) w^4}. $$ Consequently, $\epsilon = `sgn(c \mp 1))$, and $$ \kappa^{(1)} = \frac{1 \pm w^2 - \sqrt{1 + 2c w^2 + w^4}} {\sqrt{2|c \mp 1|}}, \quad \kappa^{(2)} = \frac{1 \pm w^2 + \sqrt{1 + 2c w^2 + w^4}} {\sqrt{2|c \mp 1|}}. $$ Furthermore, $$ G^{(1)} = -\frac{(1 \mp w^2 + \sqrt{1 + 2 c w^2 + w^4})^2} {2 (c \mp 1)w^2}, \quad G^{(2)} = -\frac{(1 \mp w^2 - \sqrt{1 + 2 c w^2 + w^4})^2} {2 (c \mp 1)w^2}. $$ Under the scaling by factor of $m$, the metric invariants $K^{(i)}$ and $\kappa^{(i)}$ become $K^{(i)}/m^2$ and $m \kappa^{(i)}$, respectively, while $G^{(i)}$ remains invariant. Formulas~\eqref{Gkappa} are then easily checked. Moreover, all three metric invariants are invariant under the offsetting~\eqref{offset}. Formulas for $G^{(i)}$ and $\kappa^{(i)}$ in the special cases are given in Table~\ref{T3} along with the sign $\epsilon$ of the Gaussian curvatures. \end{proof} \begin{table} $$ \begin{array}{lrcccc} & \epsilon & \kappa^{(1)} & \kappa^{(2)} & G^{(1)} & G^{(2)} \\ \hline 1. & 1 & \hphantom{-} 2\,|w| & \hphantom{-} 2\,|w| & -1 & -1 \\ \text{2a}. & 1 & \hphantom{-} \left|\frac 1{w^2} - 1\right| & \hphantom{-} |w^2 - 1| & -\frac 1{w^2} & -w^2 \\ \text{2b}. & -1 & \hphantom{-} \frac 1{w^2} + 1 & \hphantom{-} w^2 + 1 & \hphantom{-}\frac 1{w^2} & \hphantom{-} w^2 \\ \text{3a}. & 1 & -1 + \cosh w & \hphantom{-} 1 + \cosh w & \hphantom{-}\frac{1 + \cosh w}{1 - \cosh w} & \hphantom{-}\frac{1 - \cosh w}{1 + \cosh w} \\ \text{3b}. & -1 & \hphantom{-} 1 - \cos w & \hphantom{-} 1 + \cos w & \hphantom{-}\frac{1 + \cos w}{1 - \cos w} & \hphantom{-}\frac{1 - \cos w}{1 + \cos w} \\ 4. & -1 & \hphantom{-} 1 & \hphantom{-} 1 & \hphantom{-} 0 & \hphantom{-} 0 \\ \text{5a}. & -1 & \hphantom{-} \tanh^2 w & \hphantom{-} 1 & \hphantom{-}\frac 1{\sinh^2 w \cosh^2 w} & \hphantom{-} 0 \\ \text{5b}. & 1 & \hphantom{-} \tan^2 w & \hphantom{-} 1 & -\frac 1{\sin^2 w \cos^2 w} & \hphantom{-} 0 \\ \text{6a}. & -1 & \hphantom{-} \coth^2 w & \hphantom{-} 1 & \hphantom{-}\frac 1{\sinh^2 w \cosh^2 w} & \hphantom{-} 0 \\ \text{6b}. & 1 & \hphantom{-} \cot^2 w & \hphantom{-} 1 & -\frac 1{\sin^2 w \cos^2 w} & \hphantom{-} 0 \end{array} $$ \caption{Special integrable cases. Metric invariants of focal surfaces in terms of $w$.} \label{T3} \end{table} \begin{table} $$ \begin{array}{lrcccc} & \epsilon & G^{(1)} G^{(2)} & G^{(1)}(\kappa^{(1)}) & G^{(2)}(\kappa^{(2)}) \\ \hline 1. & 1 & -1 & -1 & -1 \\ \text{2a}. & 1 & \hphantom{-} 1 & -1 \pm \kappa^{(1)} & -1 \pm \kappa^{(2)} \\ \text{2b}. & -1 & \hphantom{-} 1 & -1 + \kappa^{(1)} & -1 + \kappa^{(2)} \\ \text{3a}. & 1 & -1 & -1 - \frac2{\kappa^{(1)}} & -1 - \frac2{\kappa^{(2)}} \\ \text{3b}. & -1 & \hphantom{-} 1 & -1 + \frac2{\kappa^{(1)}} & -1 + \frac2{\kappa^{(2)}} \\ 4. & -1 & \hphantom{-} 0 & \hphantom{-} 0 & \hphantom{-} 0 \\ \text{5a}. & -1 & \hphantom{-} 0 & \hphantom{-} (\sqrt{\kappa^{(1)}} - \frac1{\sqrt{\kappa^{(1)}}})^2 & \hphantom{-} 0 \\ \text{5b}. & 1 & \hphantom{-} 0 & -(\sqrt{\kappa^{(1)}} + \frac1{\sqrt{\kappa^{(1)}}})^2 & \hphantom{-} 0 \\ \text{6a}. & -1 & \hphantom{-} 0 & \hphantom{-} (\sqrt{\kappa^{(1)}} - \frac1{\sqrt{\kappa^{(1)}}})^2 & \hphantom{-} 0 \\ \text{6b}. & 1 & \hphantom{-} 0 & -(\sqrt{\kappa^{(1)}} + \frac1{\sqrt{\kappa^{(1)}}})^2 & \hphantom{-} 0 \end{array} $$ \caption{Special integrable cases. Relations between metric invariants of focal surfaces.} \label{T4} \end{table} Summarizing, focal surfaces of integrable Weingarten surfaces belong to the isometry classes specified in Proposition~\ref{prop:focal}. A natural question is whether the condition $G^{(1)} G^{(2)} = `const$ or, equivalently, $\gamma^{(1)} \gamma^{(2)} = `const$, is not only necessary, but also sufficient for the condition~\eqref{ODE} to hold. \begin{proposition} \label{prop:GG} Under the condition $\gamma^{(1)} + \gamma^{(2)} \ne 0$, a surface satisfies the governing equation~\eqref{ODE} if and only if the product $$ \numbered\label{GG} \gamma^{(1)} \gamma^{(2)} = \pm\left\|`grad^{(1)}\kappa^{(1)}\right\|^{(1)}\, \left\|`grad^{(2)}\kappa^{(2)}\right\|^{(2)} $$ is constant. \end{proposition} \begin{proof} Assuming the $\rho(\sigma)$ dependence, $\gamma^{(1)} + \gamma^{(2)}$ simplifies to $(\rho - \sigma) \rho''/\sqrt{|\rho'|^3}$ and the product in question to $$ \gamma^{(1)} \gamma^{(2)} = \frac{(\rho' - 1)^2}{\epsilon\rho'} - \frac{(\rho - \sigma)^2 \rho^{\prime\prime 2}}{4\,\epsilon\rho^{\prime3}}. $$ Factorizing the $\sigma$-derivative of this expression as $$ \pm\frac{(\rho - \sigma)^2}{2\,\epsilon\rho^{\prime3}} (\rho''' - \frac{3}{2 \rho'} \rho''{}^2 + \frac{\rho' - 1}{\rho - \sigma} \rho'' - 2 \frac{(\rho' - 1) \rho' (\rho' + 1)}{(\rho - \sigma)^2}) \rho'' $$ and comparing to the governing equation~\eqref{ODE} proves the proposition. \end{proof} It follows from the proof that condition~\eqref{GG} also holds when $\rho'' = 0$, i.e., if there is a linear relation between the radii of curvature. As of now, there seems to be no indication towards integrability of the latter class (except when $\rho \pm \sigma = `const$, which satisfies~\eqref{ODE} as well). \section{Conclusions and future work} In this work we singled out a class of Weingarten surfaces on the basis of its solitonic integrability. Although special cases were not unknown to nineteenth century geometers, the overall result appears to be new. We also characterized integrability in terms of metric invariants of the focal surfaces. For time reasons, many questions had to be left for further research. We do not know the B\"acklund transformation, recursion operator, bi-Hamiltonian structure and other attributes of integrability. We did not provide any solutions to the Gauss equation~\eqref{Gauss:gen}. We do not know what is the true geometric meaning of the spectral parameter. Even the task of computing third order symmetries of the Gauss equation proved to be too complex. We have seen in Part~I that integrability of surfaces of constant astigmatism is attributable to the fact that their focal surfaces are pseudospherical. In the general case, the existence of an integrability-preserving relation to previously known integrable surfaces is an open problem. Our nearest goals include exploring the induced Bianchi type transformation between surfaces satisfying relations~\eqref{Gkappa} as well as investigating the extended symmetries of the class in the sense of Cie\'sli\'nski~\cite{C-nls,C-G-S}. \ack We are indebted to J.~Cie\'sli\'nski, E.~Ferapontov, A.~Sergyeyev, and S.~Verpoort for advice and encouragement. The first-named author was supported by GA\v{C}R under project 201/07/P224; the second-named author by M\v{S}MT under project MSM~4781305904 ``Topologick\'e a analytick\'e metody v teorii dynamick\'ych syst\'em\accent23u a matematick\'e fyzice.'' Thanks are also due to CESNET for granting access to the MetaCentrum computing facilities. \section*{References}
\section{Introduction} \quad Let $(N,\ g)$ be an $n$-dimensional compact smooth Riemannian manifold with smooth boundary $\partial N$ and $\Delta_N$ the positive Dirichlet Laplacian on $N$. Let $L^2(N)$ be the space of square integrable functions on $N$ with respect to the Riemannian density $dv(N)=\sqrt{{\bf g}(x)}\,dx:= \sqrt{\det\ (g_{ij})}\ dx$. Let $e_1(x),\,e_2(x),\,\cdots$ be a complete orthonormal basis in $L^2(N)$ for the Dirichlet eigenfunctions of $\Delta_N$ such that $0<\lambda_1^2< \lambda_2^2\leq \lambda_3^2\leq\,\cdots$ for the corresponding eigenvalues, where $e_j(x)$ ($j=1,2,\dots$) are real valued smooth function on $N$ and $\lambda_j$ are positive numbers. Also, let ${\bf e}_j$ denote the projection of $L^2(N)$ onto the 1-dimensional space ${\bf C}e_j$. Thus , an $L^2$ function $f$ can be written as $f=\sum_{j=0}^\infty {\bf e}_j(f)$, where the partial sum converges in the $L^2$ norm. Let $\lambda$ be a positive real number $\geq 1$. We define the spectral function and the unit band spectral projection operator $\chi_\lambda$ as follows: \[e(x,y,\lambda):=\sum_{\lambda_j\leq \lambda}\,e_j(x)\,e_j(y),\] \[{\chi}_\lambda f:=\sum_{\lambda_j\in (\lambda,\,\lambda+1]}{\bf e}_j(f)\ .\] Grieser \cite{Gr} and Sogge \cite{S3} proved the $L^\infty$ estimate of $\chi_\lambda$, \begin{equation} \label{equ:sup} ||{ \chi}_\lambda f||_\infty\leq C\lambda^{(n-1)/2} ||f||_2\, \end{equation} where $||f||_r$ ($1\leq r\leq \infty$) means the $L^r$ norm of the function $f$ on $N$. In the whole of this paper $C$ denotes a positive constant which depends only on $N$ and may take different values at different places, if there is no otherwise stated. The idea of Grieser and Sogge is to use the standard wave kernel method outside a boundary layer of width $C \lambda^{-1}$ and a maximum principle argument inside that layer. By using the maximum principle argument and the estimate (\ref{equ:sup}), Xu \cite{X2} proved the gradient estimate of $\chi_\lambda$ \begin{equation} \label{equ:2infgrad} ||\nabla\, { \chi}_\lambda f||_\infty\leq C\lambda^{(n+1)/2} ||f||_2\ . \end{equation} Here $\nabla$ is the Levi-Civita connection on $N$. In particular, $\nabla f=\sum_j\, g^{ij}\partial f/\partial x_j$ is the gradient vector field of a $C^1$ function $f$, the square of whose length equals $\sum_{i,j} g^{ij}(\partial f/\partial x_i)(\partial f/\partial x_j)$. One of his motivation is to prove the H{\" o}rmander multiplier theorem on compact manifolds with boundary. Seeger and Sogge \cite{SS} firstly proved that theorem by using the parametrix of the wave kernel on manifolds without boundary. All the results mentioned in the introduction have their analog on compact manifolds without boundary. See the details in the introduction of Shi-Xu \cite{SX} and the references therein. In general, the method used in manifolds without boundary is not valid for the problems on manifolds with boundary. In particular, on manifolds with boundary the H{\" o}rmander multiplier theorem cannot be obtained by the standard pseudo-differential operator calculus as done on manifolds without boundary, since the square root of the Dirichlet Laplaican is not a pseudo-differential operator any more and one cannot obtain the $L^\infty$ bounds for $\chi_\lambda$ and $\nabla \chi_\lambda$ only by using the Hadamard parametrix of the wave kernel. In the paper, by rescaling $\chi_\lambda f$ at the scale of $\lambda^{-1}$ both outside and inside the boundary layer of width $C\lambda^{-1}$, we obtain by elliptic apriori estimates a slightly stronger estimate than (\ref{equ:2infgrad}) as follows:\\ \noindent {\bf Theorem 1.1.} {\it Let $f$ be a square integrable function on $N$. Then, for every $\lambda\geq 1$, there holds \begin{equation} \label{equ:grad2} \|\nabla \chi_\lambda f\|_\infty\leq C\left(\lambda \|\chi_\lambda f\|_\infty+\lambda^{-1}\|\Delta_N\,\chi_\lambda f\|_\infty\right). \end{equation}} \noindent {\bf Remark 1.1.} Putting $f(\cdot)=\sum_{\lambda_j\in (\lambda,\,\lambda+1]}\, e_j(x)e_j(\cdot)$ in (\ref{equ:sup}), we obtain the uniform estimate of eigenfunctions for all $x\in N$, \[\sum_{\lambda_j\in (\lambda,\,\lambda+1]}\, |e_j(x)|^2\leq C\lambda^{n-1}.\] Actually it can be proved that those two estimates are equivalent. By using the Cauchy-Scahwarz inequality, we obtain the gradient estimate (\ref{equ:2infgrad}) from Theorem 1 together with the above inequality. Similarly, that estimate is equivalent to the uniform estimate for all $x\in N$, \begin{equation} \label{equ:gradpointwise} \sum_{\lambda_j\in (\lambda,\,\lambda+1]}\, |\nabla e_j(x)|^2\leq C\lambda^{n-1}.\end{equation} \noindent{\bf Remark 1.2.} By the finite propagation speed of the wave equation, the asymptotic formula of derivatives of the spectral function $e(x,y,\lambda)$ in Theorem 1 \cite{XuB} which is proved by the standard wave kernel method, also holds for each interior point $x$ of $N$. Much more general asymptotic formulae are given in Theorems 1.8.5 and 1.8.7 of Safarov-Vassiliev \cite{SV}. In particular, we have the following asymptotic formula that as $\lambda\to\infty$ \[\sum_{\lambda_j\leq \lambda}|\nabla\,e_j(x)|^2=\frac{n\,\lambda^{n+2}}{2\,(4\pi)^{n/2}\,\Gamma(2+\frac{n}{2})}+{\rm O}_x(\lambda^{n+1}),\] where the constant in the reminder term ${\rm O}_x(\lambda^{n+1})$ depends on the distance of $x$ to the boundary of $N$. Hence, the exponents of $\lambda$ in estimates (\ref{equ:2infgrad}) and (\ref{equ:gradpointwise}) are sharp at $x$ as $\lambda\to\infty$. For each point $z$ on the boundary of $N$, Ozawa \cite{Oz} used the heat kernel method to show the asymptotic formula that as $\lambda\to\infty$, \[\sum_{\lambda_j\leq \lambda}\left|\frac{\partial e_j}{\partial \nu}(z)\right|^2=\frac{\lambda^{n+2}}{(4\pi)^{n/2}\,\Gamma(2+\frac{n}{2})}+{\rm o}(\lambda^{n+2}),\] where $\nu$ is the unit outer normal vector field on the boundary of $N$. Hence, (\ref{equ:2infgrad}) and (\ref{equ:gradpointwise}) are also sharp on the boundary. \\ \noindent {\bf Corollary 1.1.} {\it Let $e_\lambda(x)$ be an eigenfunction with respect to the positive Dirichlet Laplacian $\Delta_N$ on $N$: $\Delta_M e_\lambda=\lambda^2 e_\lambda$ in the interior of $N$ and $e_\lambda=0$ on the boundary of $N$. Then, for every $\lambda\geq 1$, there holds the upper bound estimate of $\nabla e_\lambda$ {\rm :} \[\|\nabla e_\lambda\|_\infty\leq C\lambda\,\|e_\lambda\|_\infty.\]} \noindent {\bf Proof. } Putting $f=e_\lambda$ in the estimate (\ref{equ:grad2}), we obtain the corollary. \hfill{$\Box$} \\ Actually, by the basic geometric property of nodal sets of an eigenfunction, we can find a complete picture for the $L^\infty$ norm of $\nabla e_\lambda$ in the following:\\ \noindent {\bf Theorem 1.2.} {\it Let $e_\lambda(x)$ be an eigenfunction with respect to the Dirichlet Laplacian operator $\Delta_N$ on $N$ without boundary: $\Delta_N e_\lambda=\lambda^2 e_\lambda$. Then, for every $\lambda\geq 1$, there holds \begin{equation} \label{equ:grad} \lambda\|e_\lambda\|_\infty/C\leq \|\nabla e_\lambda\|_\infty\leq C\lambda \|e_\lambda\|_\infty. \end{equation} }\\ \noindent {\bf Remark 1.3.} The authors \cite{SX} proved the analog of Theorems 1 and 2 on compact manifolds without boundary. The proof in this paper is more complicated than there because we need to do analysis at points near the boundary. We believe that there also hold the analog for $k$-covariant derivatives $\nabla^k\,\chi_\lambda f$ and $\nabla^k \,e_\lambda$ on $N$. We plan to discuss this question in a future paper.\\ \noindent {\bf Remark 1.4.} Let $\psi_j$ be the normal derivative of $e_j$ at the boundary $\partial N$ of $N$. The lower bound estimate \[\|\psi_j\|_{L^\infty(\partial N)}\geq C\|e_j\|_{L^\infty(N)}\] does not hold in general. Using Examples 3-5 in Hassell-Tao \cite{HT} and doing a little bit more computations, we can see that the above estimate does not hold on the flat cylinder, the hemisphere and the spherical cylinder. We hope to find a sufficient condition for the lower bound estimate in a future work.\\ We conclude the introduction by explaining the organization of this paper. In Section 2, we show the lower bound of the gradient $\nabla\, e_\lambda$ by the basic geometrical property of the nodal set of eigenfunctions. In Section 3 we use the rescaling method and the H{\" o}lder estimate about elliptic PDEs to show (\ref{equ:grad2}) and the upper bound part of (\ref{equ:grad}). The point is to do the rescaling both outside and inside the boundary layer of width $C\,\lambda^{-1}$. \section{Lower bound of $\nabla e_\lambda$} \quad \ \ The nodal set of an eigenfunction $e_\lambda$ of $\Delta_N$ is the zero set \[Z_{e_\lambda}:=\{x\in N: e_\lambda(x)=0\}.\] A connected component of the open set $N\backslash Z_{e_\lambda}$ is called a nodal domain of the eigenfunction $e_\lambda$. We have the same definition for manifolds without boundary.\\ \noindent {\bf Lemma 2.1.} (Br{\" u}ning \cite{Br}) {\it Let $M$ be a compact Riemannian manifold without boundary. Let $\lambda\geq 1$ and $e_\lambda$ be an eigenfunction of the positive Laplacican $\Delta_M$: $\Delta_M\,e_\lambda=\lambda^2 e_\lambda$. Then there exists a constant $C$ only depending on $M$ such that each geodesic ball of radius $C/\lambda$ in $M$ must intersect the nodal set $Z_{e_\lambda}$ of $e_\lambda$.} \noindent{\bf Proof. } A proof written in English is given by Zelditch in pp. 579-580 of \cite{Z}.\hfill{$\Box$}\\ We need a manifold-with-boundary version of Lemma 1 as follows:\\ \noindent {\bf Lemma 2.2.} {\it Let $\lambda>0$ and $\lambda^2$ be greater than the smallest eigenvalue $\lambda_1^2$ of the Dirichlet Laplacian $\Delta_N$. Let $e_\lambda$ be an eigenfunction of $\Delta_N$: $\Delta_N e_\lambda=\lambda^2 e_\lambda$ in the interior ${\rm Int}(N)$ of $N$ and $e_\lambda=0$ on the boundary $\partial N$ of $N$. Then there exists a positive constant $D$ only depending on $N$ such that each geodesic ball of radius $D/\lambda$ contained in ${\rm Int}(N)$ must intersect the nodal set $Z_{e_\lambda}$ of $e_\lambda$.}\\ \noindent {\bf Proof. } We here adapt the proof of Zelditch \cite{Z} with a slight modification.\\ {\it Step 1}\quad We show the following fact: There exists a constant $C$ such that for each interior point $p$ of $N$ and each positive number $r>0$ satisfying that the distance $d(p,\,\partial N)<r$ from $p$ to $\partial N$ is less than $r$, the smallest Dirichlet eigenvalue $\lambda_1^2\bigl(B(p,\,r)\bigr)$ of the geodesic ball $B(x,\,r)$ is bounded from above by $C/r^2$. Since $d(p,\,\partial N)>r$, we may assume that there exists a geodesic normal coordinate chart $(x_1,\dots, x_n)$ on the ball $B(p,\,r)$. Let $g$ be the Riemannian metric of $N$ on $B(p,\,r)$ with coefficients $g_{ij}=g(\partial/\partial x_i,\,\partial/\partial x_j)$ and $g_0$ be the Euclidean metric $dx_1^2+\cdots+dx_n^2$ on $B(x,\,r)$. Take $0<c_1<1$ depending only on $N$ so that the Euclidean ball $B(p,c_1r\,;\,g_0)$ is contained in the metric ball $B(p,r\,;\,g)$ of $N$. Since $c_1<1$, by the definition of Rayleigh quotient, \[\lambda_1^2\bigl(B(p,r\,;\,g)\bigr)\leq \lambda_1^2\bigl(B(p,c_1r\,;\,g)\bigr).\] Since $N$ is compact, by comparing Rayleigh quotients, there exists $c_2>0$ depending only on $N$ such that \[\lambda_1^2\bigl(B(p,c_1r\,;\,g)\bigr)\leq c_2 \lambda_1^2\bigl(B(p,c_1r\,;\,g_0)\bigr).\] On the other hand, by change of variables, we have \[\lambda_1^2\bigl(B(p,c_1r\,;\,g_0)\bigr)=\frac{\lambda_1^2\bigl(B(p,1\,;\,g_0)\bigr)}{(c_1r)^2}.\] Combining the above three inequalities and setting $C=c_2c_1^{-2}\lambda_1^2\bigl(B(p,1\,;\,g_0)\bigr)$, we complete the proof.\\ {\it Step 2} Take a geodesic ball $B(p,\,r)$ in $N$ such that $d(p,\,\partial N)>r$. Suppose that it is disjoint from the nodal set $Z_{e_\lambda}$. Then it is completely contained in a nodal domain $D_j$ of $e_\lambda$. But $\lambda^2=\lambda_1^2(D_j)\leq \lambda_1^2\bigl(B(p,\,r)\bigr)\leq C/r^2$. Hence, $r\leq \sqrt{C}{\lambda}$. Taking $D=2\sqrt{C}$, we complete the proof.\hfill{$\Box$} \\ \noindent {\sc Proof of the lower bound part of Theorem 2}\quad Take a point $x$ in $N$ such that $|e_\lambda(x)|=\|e_\lambda\|_\infty$. By the Dirichlet boundary condition, the distance $d$ from $x$ to $\partial N$ is positive. {\it Case 1} \quad Assume $d>D/\lambda$. Then there exists point $y$ in the geodesic ball $B(x,\,D/\lambda)$ with center $x$ and radius $D/\lambda$ such that $e_\lambda(y)=0$. We may assume $\lambda$ so large that there exists a geodesic normal chart $(r, \theta)\in [0,\,D/\lambda]\times {\Bbb S}^{n-1}(1)$ in the ball $B(x,\,D/\lambda)$. By the mean value theorem, there exists a point $z$ on the geodesic segment connecting $x$ and $y$ such that \[\left|\frac{\partial e_\lambda}{\partial r}(z)\right|\geq \frac{\lambda}{D}|e_\lambda(x)|=\frac{\lambda}{D}\|e_\lambda\|_\infty.\] {\it Case 2} \quad Assume $d\leq D/\lambda$. We may assume $\lambda$ so large that there exists a unique geodesic $\gamma:[0,\,d]\to N$ of arc length parameter connecting $x$ and $\partial N$, \[\gamma(0)=x,\quad \gamma(d)\in \partial N.\] Since $e_\lambda(\gamma(d))=0$, by the mean value theorem, there exists $t_0$ in $(0,\,d)$ such that \[\left|\frac{d e_\lambda\bigl(\gamma(t)\bigr)}{dt}(t_0)\right|=\frac{|e_\lambda(x)|}{d}\geq \frac{\lambda}{D}\,\|e_\lambda\|_\infty.\] \hfill{$\Box$} \section{Estimate for $\nabla \chi_\lambda f$} \subsection{Outside the boundary layer} \quad\ \ Recall the principle: {\it On a small scale comparable to the wavelength $1/\lambda$, the eigenfunction $e_\lambda$ behaves like a harmonic function.} It was developed in H. Donnelly and C. Fefferman \cite{DF1} \cite{DF2} and N. S. Nadirashvili \cite{Na} and was used extensively there. Recently Mangoubi \cite{Ma1} applied this principle to studying the geometry of nodal domains of eigenfunctions. In this section, for a square integrable function $f$ on $N$ we give a modification of this principle, which can be applied to the Poisson equation \[\Delta_N\,\chi_\lambda f=\sum_{\lambda_j\in (\lambda,\,\lambda+1]}\, \lambda_j^2 {\bf e}_j(f)\quad {\rm in}\quad {\rm Int}(N)\] with the Dirichlet boundary condition $\chi_\lambda f=0$ on $\partial N$. In particular, in this subsection, we do the analysis outside the boundary layer $L_{1/\lambda}=\{z\in N:\,d(z,\,\partial N)\leq 1/\lambda\}$ of width $1/\lambda$. Take an arbitrary point $p$ with $d(p,\,\partial N)\geq 1/\lambda$. We may assume that $1/\lambda$ is sufficiently small such that there exists a geodesic normal coordinate chart $(x_1,\dots, x_n)$ on the geodesic ball $B(p,\,2/\lambda)$ in $N$. In this chart, we may identify the ball $B(p,\,2/\lambda)$ with the $n$-dimensional Euclidean ball ${\Bbb B}(2/\lambda)$ centered at the origin $0$, and think of the function $\chi_\lambda f$ in $B(p,\,2/\lambda)$ as a function in ${\Bbb B}(2/\lambda)$. Our aim in this subsection is to show the inequality \begin{equation} \label{equ:outside2} |(\nabla \chi_\lambda f)(p)|\leq C\left(\lambda\|\chi_\lambda f\|_{L^\infty\bigl({\Bbb B}(2/\lambda)\bigr)}+\lambda^{-1}\|\Delta_N\,\chi_\lambda f\|_{L^\infty\bigl({\Bbb B}(2/\lambda)\bigr)}\right). \end{equation} For simplicity of notions, we rewrite $u=\chi_\lambda f$ and $v=\Delta_N\chi_\lambda f$ in what follows. The Poisson equation satisfied by $u$ in ${\Bbb B}(1/\lambda)$ can be written as \[-\frac{1}{\sqrt{g}}\sum_{i,j}\,\partial_{x_i}\left(g^{ij}\sqrt{g}\partial_{x_j} u\right)=v.\] Consider the rescaled functions $u_\lambda(y)=u(y/\lambda)$ and $v_\lambda(y)=v(y/\lambda)$ in the ball ${\Bbb B}(2)$. The above estimate we want to prove is equivalent to its rescaled version \begin{equation} \label{equ:scale} |(\nabla u_\lambda)(0)|\leq C\left(\|u_\lambda\|_{L^\infty\bigl({\Bbb B}(2)\bigr)}+\lambda^{-2}\|v_\lambda\|_{L^\infty\bigl({\Bbb B}(2)\bigr)}\right). \end{equation} On the other hand, the rescaled version of the Poisson equation has the expression, \begin{equation} \label{equ:PoissonScaled} \sum_{i,j}\,\partial_{y_i}\left(g^{ij}_\lambda\sqrt{g_\lambda}\partial_{y_j} u_\lambda\right)=-\lambda^{-2}\,\sqrt{g_\lambda}\,v_\lambda,\end{equation} where $g_{ij,\,\lambda}(y)=g_{ij}(y/\lambda)$, $g^{ij}_\lambda(y)=g^{ij}(y/\lambda)$ and $\sqrt{g_r}(y)=(\sqrt{g})(y/\lambda)$. For each $0<\alpha<1$, there exists $K>0$ such that the $C^\alpha$ norm of the coefficients $g^{ij}_\lambda\sqrt{g_\lambda}$, $\sqrt{g_\lambda}$ in ${\Bbb B}(2)$ are bounded uniformly from above by $K$, and the smallest eigenvalue of the $n\times n$ matrix $(g^{ij}_\lambda\sqrt{g_\lambda})_{ij}$ in ${\Bbb B}(2)$ bounds from below by $1/K$, for all $\lambda\geq 1$. By Theorem 8.32 in page 210 of Gilbarg-Trudinger \cite{GT}, there exists constant $C=C(n, \, \alpha,\, K)$ such that \[\|u_\lambda\|_{C^{1,\,\alpha}\bigl({\Bbb B}(1)\bigr)}\leq C\left( \|u_\lambda\|_{L^\infty\bigl({\Bbb B}(2)\bigr)}+\lambda^{-2}\,\|v_\lambda\|_{L^\infty\bigl({\Bbb B}(2)\bigr)}\right), \] This is stronger than the estimate (\ref{equ:scale}). Therefore, we complete the proof of Theorem 1.1 outside the boundary layer $L_\lambda$. \subsection{Inside the boundary layer} \quad\ \ Using the notions in subsection 3.1, We are going to prove the following estimate: \begin{equation} \label{equ:inside} \|\nabla u\|_{L^\infty(L_{1/\lambda})}\leq C\left(\lambda \|u\|_\infty+\lambda^{-1}\|v\|_\infty\right), \end{equation} with which combining (\ref{equ:outside2}) completes the proof of Theorem 1.1. We may assume that $\lambda$ is sufficiently large so that there exists a geodesic normal coordinate chart $(z',\,z_n)$ on the boundary layer $L_{3/\lambda}=\{p\in N:\,d(p,\,\partial N)\leq 3/\lambda\}$ with respect to the boundary $\partial N$. Hence, for each point $(z',\,z_n)\in L_{3/\lambda}$, we have $0\leq z_n\leq 3/\lambda$ and \[d\bigl((z',\,z_n),\,\partial N\bigr)=z_n.\] For each point $q\in \partial N$ and $r>0$, denote by $B_+(q,\,r)$ the set of points of $N$ with distance less than $r$ to $q$. Denote by ${\Bbb B}_+(r)$ the upper half Euclidean ball \[\{x=(x_1,\dots,x_n)\in {\bf R}^n:|x|<r,\,x_n\geq 0\}\] centered at the origin and with radius $r$. Then, for each $q\in \partial N$, there exists a geodesic normal chart on $B_+(q,\,3/\lambda)$ such that the exponential map $\exp_q$ at $q$ gives a diffeomorphism from ${\Bbb B}_+(3/\lambda)$ onto $B_+(q,\,3/\lambda)$. Since $\{B_+(q,\,2/\lambda):q\in \partial N\}$ forms an open cover of $L_{1/\lambda}$, the question can be reduced to showing the analog of (\ref{equ:inside}) on $B_+(q,\,2/\lambda)$ for each $q$. We only need to prove its equivalent rescaled version, \begin{equation} \label{equ:insidescaled} \|\nabla u_\lambda\|_{L^\infty\bigl({\Bbb B}_+(2)\bigr)}\leq C\left( \|u_\lambda\|_{L^\infty\bigl({\Bbb B}_+(3)\bigr)}+\lambda^{-2}\|v_\lambda\|_{L^\infty\bigl({\Bbb B}_+(3)\bigr)}\right), \end{equation} where $u_\lambda$ and $v_\lambda$ are the the rescaling function of $u$ an $v$, respectively. Observe that $u_\lambda$ and $v_\lambda$ satisfy the Poisson equation $(\ref{equ:PoissonScaled})$ in the upper half Euclidean ball ${\Bbb B}(3)$ and the Dirichlet boundary condition, \[u_\lambda=0\quad {\rm on\ the\ portion}\quad \{x\in {\Bbb B}(3):x_n=0\}\] of the boundary of ${\Bbb B}_+(3)$. For each $0<\alpha<1$, there exists $K>0$ such that the $C^\alpha$ norm of the coefficients $g^{ij}_\lambda\sqrt{g_\lambda}$, $\sqrt{g_\lambda}$ in ${\Bbb B}_+(3)$ are bounded uniformly from above by $K$, and the smallest eigenvalue of the $n\times n$ matrix $(g^{ij}_\lambda\sqrt{g_\lambda})_{ij}$ in ${\Bbb B}_+(3)$ bounds from below by $1/K$, for all $\lambda\geq 1$. By Theorem 8.36 in page 212 of Gilbarg-Trudinger \cite{GT}, there exists constant $C=C(n, \, \alpha,\, K)$ such that \[\|u_\lambda\|_{C^{1,\,\alpha}\bigl({\Bbb B}_+(2)\bigr)}\leq C\left( \|u_\lambda\|_{L^\infty\bigl({\Bbb B}_+(3)\bigr)}+\lambda^{-2}\,\|v_\lambda\|_{L^\infty\bigl({\Bbb B}_+(3)\bigr)}\right). \] This is a stronger estimate than (\ref{equ:insidescaled}). \\ \noindent {\bf Acknowledgements}\quad Yiqian Shi is supported in part by the National Natural Science Foundation of China (No. 10671096 and No. 10971104) and Bin Xu by the National Natural Science Foundation of China (No. 10601053 and No. 10871184).
\section{Introduction} In the recent years, the mathematical study of the thermistor problem has been of interest by various authors (see for instance \cite{ac,cim,com,mg,xu} and the references therein) however the cross effects are neglected. Here we study a mathematical model for thermoelectric conductors, introduced in \cite{zamm}, taking into account the presence of the Seebeck-Peltier-Thomson and the Joule effects. Indeed, in the thermodynamics analysis, the Joule effect is given by $|{\bf j}|^2/ \sigma(\cdot,\theta)$ with (cf. \cite{j}) \begin{equation} {\bf j}=- \sigma(\cdot,\theta)(\alpha(\cdot,\theta)\nabla\theta+\nabla\phi),\label{dj} \end{equation} representing the current density, $\theta$ denoting the temperature, $\phi$ is the electric potential, and the electrical conductivity $\sigma$ is a known positive function. The electrical conductivity is assumed temperature dependent and this is different at different places along the material due to the molecular structure. The Seebeck coefficient $\alpha$ is a given nonlinear function, dependent both in space and temperature, with constant sign observing that the sign of the Seebeck coefficient corresponds to the sign of the Hall effect \cite{ioffe}. The Thomson effect is $-\partial\alpha/\partial T(\cdot,\theta)\theta\nabla\theta\cdot{\bf j}$, where $\partial/\partial T$ means the derivative with respect to the real variable. Due to the first Kelvin relation, $\pi(\theta)=\theta\alpha(\theta)$, $\theta\alpha(\theta)$ corresponds to the Peltier coefficient. Due to the second Kelvin relation, $\mu(\theta)=\theta\alpha'(\theta)$, this coefficient is known as Thomson coefficient. Although $\mu$ is the only thermoelectric coefficient directly measurable for individual materials \cite{gasser,km}, and the Seebeck coefficient appears as $\alpha(T)=\int_{T_r} ^T {\mu(t)\over t}dt$ for some reference temperature $T_r$, we keep the Seebeck coefficient as a given function as it is usual in the literature for the thermistor problem when this cross-effect coefficient is taken into account \cite{badii,cim2,xu2}. Thus, the resulting PDE's system is strongly coupled. The thermoelectric problem under study reads (see its derivation in \cite{zamm,j}) \noindent ($\mathcal P$) Find the pair temperature-potential $(\theta,\phi)$ such that \begin{eqnarray} -\nabla\cdot(k\nabla\theta)= \sigma(\cdot,\theta)\alpha(\cdot,\theta)(\alpha(\cdot,\theta)+{\partial\alpha\over\partial T}(\cdot,\theta)\theta)|\nabla\theta|^2 +\nonumber\\+ \sigma(\cdot,\theta)(2\alpha(\cdot,\theta)+{\partial\alpha\over\partial T}(\cdot,\theta)\theta)\nabla\theta\cdot \nabla\phi+ \sigma(\cdot,\theta)|\nabla\phi|^2+g &&\mbox{ in }\Omega,\label{pbu}\\ \label{defphi} -\nabla\cdot(\sigma(\cdot,\theta)\nabla\phi)=\nabla\cdot \left({ \sigma(\cdot,\theta)}\alpha(\cdot,\theta)\nabla\theta\right) &&\mbox{ in }\Omega, \\ k\nabla \theta\cdot{\bf n}=-\alpha(\cdot,\theta)\theta h &&\mbox{ on }\partial\Omega, \label{kh}\\ \sigma(\cdot,\theta)(\nabla\phi+\alpha(\cdot,\theta)\nabla\theta)\cdot{\bf n}=h &&\mbox{ on }\partial\Omega,\label{pbn} \end{eqnarray} where $\Omega $ is a convex bounded domain of $\mathbb R^n$ $(n\geq 2)$, and it may represent electrically conductive rigid solids such as for instance thermistors, thermocouples, resistive thermal devices (also called resistance temperature detectors) or thermoelectric coolers (see \cite{bulu,chakra,yama} and the references therein). Notice that the boundary $\partial\Omega$ is Lipschitz since every bounded convex open subset of $\mathbb R^n$ always has a Lipschitz boundary \cite[Section 1.2]{grisv}. Here $\bf n$ is the unit outward normal to the boundary $\partial\Omega$, $g$ denotes the external heat sources and $h$ denotes the surface current source. The thermal conductivity $k$ is the known positive coefficient of the Fourier law. Finally, we remark that (\ref{kh}) is known as the linear Newton law of cooling with $\alpha(\theta)h$ representing the heat transfer coefficient. Our two main purposes are: (1) to improve the existence result stated in \cite{zamm} for three dimensional space, i.e. the solution belongs to the Sobolev space $W^{1,p}(\Omega)$ with $p>n=3$. Indeed, we state the existence result for the multidimensional case since its proof is valid for any $n\geq 2$; (2) to pass to the limit on the thermal conductivity, $k\rightarrow+\infty$, in order to show the existence of the limit model \begin{eqnarray}\label{t1} \nabla\cdot (\sigma(\cdot,\Theta)\nabla\phi)=0 \quad\mbox{ in }\Omega;\\ \sigma(\cdot,\Theta)\nabla\phi\cdot{\bf n}=h \quad\mbox{ on }\partial\Omega,\label{t2} \end{eqnarray} for some positive constant $\Theta$ solving an implicit scalar equation. The nonlocal problem (\ref{t1})-(\ref{t2}) is known as the shadow system, the heuristic designation introduced by Nishiura \cite{n} in order to exhibit minimal dynamics displaying the mechanism of basic pattern formation. Here the shadow system turns out the dynamic relation among the trivial rest states (constant solutions) and the large amplitude voltages. The present result generalizes the homogeneous Neumann boundary value problem already studied for nonlocal elliptic problems \cite{cr}. The regularity assumptions about the domain could be weakened if different techniques are provided (see \cite{bw,dk,dauge,elsch,med,zan} and the references therein). We refer to \cite{costa} the existence of singularities of electromagnetic fields at corners and edges of a bounded Lipschitz domain with piecewise plane boundary. The contents of this work are as follow. In next Section we state the assumptions and the main results. Section \ref{aux} deals with existence and regularity results for auxiliary problems. We prove in Section \ref{sexist} the existence result for $n\geq 2$ when the electrical conductivity is assumed be a uniformly continuous function, and we obtain in Section \ref{sexist1} its limit model. The proof of the existence result valid for $n= 2$ when the electrical conductivity is assumed be discontinuous on the space variable and the corresponding asymptotic limit model are postponed in Section \ref{2d}. \section{Assumptions and main results} In order to establish the existence results we assume the following set of hypotheses on the data. \begin{description} \item[(H1)] $k$ is a positive constant. \item[(H2)] $\sigma:\Omega\times\mathbb R\rightarrow\mathbb R$ is a Carath\'eodory function, i.e. measurable with respect to $x\in\Omega$ and continuous with respect to $T\in\mathbb R$, and furthermore \begin{equation} \exists \sigma_\#,\sigma^\#>0:\quad \sigma_\#\leq \sigma(\cdot,T)\leq \sigma^\#,\quad\forall T\in \mathbb R,\ \mbox{a.e. in }\Omega.\label{smin}\end{equation} \item[(H3)] $ g$ belongs to the Lebesgue space $ L^{p/2}(\Omega)$, $p>2$, and $h\in C(\partial\Omega)$ is such that verifies the compatibility condition \begin{equation}\label{cc} \int_{\partial\Omega}h\ ds=0, \end{equation} where $ds$ represents the element of surface area. \item[(H4)] $\alpha\in C^{}(\bar\Omega\times\mathbb R)$ is such that \begin{equation} \exists L_\alpha>0:\quad|{\alpha}(x_1,T_1)-{\alpha}(x_2,T_2)|\leq L_\alpha(|x_1-x_2|+|T_1-T_2|),\quad \label{la} \end{equation} for all $x_1,x_2\in\bar\Omega$ and for all $T_1,T_2\in \mathbb R$, with $|x|$ representing the euclidean norm and $|T|$ the absolute value of a real number. Moreover, for all $x\in\Omega$, the mapping $T\mapsto\alpha(x,T)$ is continuously differentiable in $\mathbb R$ and its derivative satisfies \begin{equation} \exists\mu^\#>0:\quad|{\partial\alpha\over\partial T}(x,T)|\leq \left\{\begin{array}{ll} \mu^\#,\quad&\mbox{ if }|T|\leq 1\\ \mu^\#/|T|,\quad&\mbox{ if }|T|>1 .\end{array}\right. \label{alphat} \end{equation} The following two different cases will be addressed: \begin{description} \item[(H4)$_+$ for materials with positive $\alpha$:] \begin{equation} \exists\alpha_\#,\alpha^\#>0:\quad \alpha_\#\leq \alpha(x,T)\leq \alpha^\#,\quad\forall x\in \bar\Omega,\quad\forall T\in \mathbb R.\label{amm} \end{equation} Moreover, there exist an open subset $\Gamma\subset \partial\Omega$ such that meas$(\Gamma)>0$ and meas$(\partial \Omega\setminus\bar\Gamma)>0$ and constants $h_\#>0$ and $h_1<0$ such that \begin{eqnarray}\label{defh}\qquad h_1>-{\min\{k,\alpha_\#h_\#\}\over C_1\alpha^\#}, \quad h(x)\geq \left\{\begin{array}{ll} h_\#, \quad&\mbox{a.e. } x\in \Gamma\\ h_1, &\mbox{a.e. }x\in\Sigma, \end{array}\right. \end{eqnarray} where $\Sigma:=\partial \Omega\setminus\bar\Gamma$ and $C_1$ denotes the continuity constant of the embedding $H^1(\Omega)\hookrightarrow L^2(\Sigma)$, i.e. \begin{equation}\label{cc1} \|\theta\|_{2,\Sigma}^2\leq C_1\left(\|\nabla\theta\|_{2,\Omega}^2+\|\theta\|_{2,\Gamma}^2\right) \end{equation} for every $\theta\in H^1(\Omega)$; \item[(H4)$_-$ for materials with negative $\alpha$:] \begin{equation} \exists\alpha_\#,\alpha^\#>0:\quad - \alpha^\#\leq\alpha(x,T)\leq -\alpha_\#,\quad\forall x\in \bar\Omega,\quad\forall T\in \mathbb R.\label{amm-} \end{equation} Moreover, there exist an open subset $\Gamma\subset \partial\Omega$ such that meas$(\Gamma)>0$ and meas$(\partial \Omega\setminus\bar\Gamma)>0$ and constants $h_\#>0$ and $h_1<0$ such that \begin{eqnarray}\label{defh-}\qquad h_1>-{\min\{k,\alpha_\#h_\#\}\over C_1\alpha^\#}, \quad h(x)\leq \left\{\begin{array}{ll} - h_\#, \quad&\mbox{a.e. } x\in \Gamma\\ -h_1, &\mbox{a.e. }x\in\Sigma. \end{array}\right. \end{eqnarray} \end{description} \end{description} \begin{remark} In particular, (\ref{cc1}) holds for the unity function then we have \begin{equation}\label{c1} |\Sigma|\leq C_1|\Gamma|, \end{equation} where $|\Gamma|=$meas$(\Gamma)$ and $|\Sigma|$=meas$(\partial \Omega\setminus\bar\Gamma)$. \end{remark} {\bf Some remarks on the assumptions.} \begin{enumerate} \item The heat conductivity can be an uniformly continuous function on both variables verifying \[\exists k_\#,k^\#>0:\quad k_\#\leq k(x,T)\leq k^\#,\quad\forall x\in \Omega,\quad\forall T\in \mathbb R. \] Since our purpose is to study the asymptotic behavior as $k\rightarrow \infty$, we assume it as constant. \item The case $\Gamma=\partial\Omega$ is excluded, from the fact that the Gauss theorem yields the necessary condition (\ref{cc}) of the existence of a solenoidal function satisfying (\ref{dj}) and (\ref{pbn}) (cf. (\ref{defphi})). \item From the assumption (\ref{alphat}) we obtain that $|{\partial\alpha\over\partial T}(x,T)|\leq \mu^\#,$ for all $x\in \Omega$ and for all $T\in\mathbb R$, and that the Thomson coefficient is bounded, i.e. $|\mu(x,T)|=|T{\partial\alpha\over\partial T}(x,T)|\leq \mu^\#$, for all $x\in \Omega$ and for all $T\in \mathbb R$. \item By the Weierstrass Theorem, any continuous function defined on a compact set (of $\mathbb R^{n+1}$) is bounded. Then the upper bound in (\ref{amm}) as well as the lower bound in (\ref{amm-}) could be given reduced to $|T|>1$. \end{enumerate} We define the Nemytskii operators \begin{eqnarray*} \sigma(\theta)=\sigma(\cdot,\theta(\cdot));\qquad \alpha(\theta)=\alpha(\cdot,\theta(\cdot));\qquad {\partial\alpha\over\partial T}(\theta)={\partial\alpha\over\partial T}(\cdot,\theta(\cdot)), \end{eqnarray*} that map $L^1(\Omega)$ into $L^q(\Omega)$, for all $q<\infty$. Their designation is kept in order to clarify the presentation. We endow the Sobolev space $W^{1,p}(\Omega)$, $p>1$, with the equivalent norms: \begin{itemize} \item for the temperature solution \[ \|\theta\|_{1,p,\Omega}=\|\nabla\theta\|_{p,\Omega}+\|\theta\|_{p,\Gamma};\] \item for the potential solution \[\|\phi\|_{1,p,\Omega}=\|\nabla\phi\|_{p,\Omega},\] considering the correspondent Poincar\'e inequality. \end{itemize} \begin{definition}\label{d1} We say that $(\theta,\phi)$ is a {\bf weak solution} to (\ref{pbu})-(\ref{pbn}) if $(\theta,\phi)\in W^{1,p}(\Omega)^2$, for $p> n$, and it satisfies \begin{eqnarray} k\int_\Omega \nabla\theta\cdot\nabla\eta dx +\int_{\partial\Omega}\alpha(\theta)h\theta \eta ds=\nonumber\\ = \int_\Omega \sigma(\theta)\Big(\alpha(\theta)(\alpha(\theta)+ {\partial\alpha\over\partial T}(\theta)\theta)|\nabla\theta|^2+(2\alpha(\theta)+{\partial\alpha\over\partial T}(\theta)\theta)\nabla\theta\cdot\nabla\phi\Big)\eta dx \nonumber\\ +\int_\Omega \left({\sigma(\theta)}|\nabla\phi |^2+g\right) \eta dx,\quad\forall \eta \in W^{1,p'}(\Omega);\qquad\label{heatj}\\ \int_\Omega\sigma(\theta)\nabla\phi\cdot \nabla\eta dx=\nonumber\\ =-\int_\Omega\sigma (\theta)\alpha(\theta)\nabla\theta\cdot\nabla\eta dx +\int_{\partial\Omega} h\eta ds,\quad\forall \eta\in W^{1,p'}(\Omega),\qquad\label{pbphij} \end{eqnarray} where $p'$ denotes the conjugate exponent to $p$, $p'=p/(p-1)$. \end{definition} \begin{remark} \label{mpnl} If $\Omega$ is of class $C^{0,1}$ and $mp>n,$ then the Morrey-Sobolev embedding holds \[W^{m,p}(\Omega)\hookrightarrow C^{0,m-n/p}(\bar\Omega).\] The vector field $\bf n$ belongs only to ${\bf L}^\infty(\partial\Omega)$ if it is the unit outward normal vector to the boundary of $C^{0,1}$ domains. When $p>{n},$ the embedding $W^{1,p'} (\Omega)\hookrightarrow L^{pn/(pn-n-p)}(\Omega)$ is valid. Thus the $L^{p/2}$ behavior of the quadratic terms $|\nabla\theta|^2$, $\nabla\theta\cdot\nabla\phi$, $|\nabla\phi|^2$ is meaningful on the right hand side of (\ref{pbu}) since $p/2>pn/(p+n)$.\end{remark} Let us extend the existence results whose can be found in \cite{zamm}. The first main theorem states the existence of weak solutions to the problem under study, strengthening the assumption (H2), i.e. strengthening the regularity on $\sigma$. \begin{theorem}\label{exist} Assume $n\geq 2$, (H1) and (H3)-(H4) hold. Additionally we assume that $ \sigma:\Omega\times\mathbb R\rightarrow\mathbb R$ is a uniformly continuous function satisfying (\ref{smin}) and the smallness condition (\ref{ss}) is satisfied. Then the variational problem (\ref{heatj})-(\ref{pbphij}) admits a weak solution $(\theta,\phi)\in W^{2,pn/(p+n)}(\Omega)\times W^{1,p}(\Omega)$, with $p>n$, in the sense of Definition \ref{d1}, such that $\int_{\partial\Omega}\phi ds$. Moreover, the following estimates hold \begin{eqnarray} \|\nabla\phi\|_{2,\Omega} &\leq& {1\over\sigma_\#}(\sigma^\#\alpha^\#\|\nabla\theta\|_{2,\Omega}+C_2\|h\|_{2,\partial\Omega} ) ; \label{cotaphi}\\ \label{cotaup} \|\theta\|_{1,p,\Omega}&\leq &C_p;\\ k\|\nabla \theta\|_{2,\Omega}^2&\leq& C_3\Big(\alpha^\#|h_1|C_p^2+\nonumber\\ \label{cotau} &&+(AC_p^2+BC_p \|\nabla\phi\|_{2,\Omega}+\sigma^\#\|\nabla\phi\|_{2,\Omega}^2+\|g\|_{p/2,\Omega})C_p\Big), \end{eqnarray} where $C_2$ denotes the Poincar\'e-Sobolev continuity constant of the embedding $H^1(\Omega)\hookrightarrow L^2(\partial\Omega)$, $C_3$ denotes the continuity constant of the embedding $W^{1,p}(\Omega)\hookrightarrow C(\bar\Omega)$, \begin{equation}\label{ab} A:=\sigma^\#\alpha^\#(\alpha^\#+\mu^\#),\qquad B:=\sigma^\#(2\alpha^\#+\mu^\#); \end{equation} and $C_p$ denotes a positive constant independent on $k$ if $k>\alpha_\#h_\#$. \end{theorem} If we assume $\sigma\in C(\bar\Omega\times\mathbb R)$ as in \cite{zamm}, the Nemytskii operator maps $C(\bar\Omega)$ into $C(\bar\Omega)$ which implies the uniform continuity on the spatial variable. Next, we establish the existence of a solution $\phi=\phi(\Theta)$ to (\ref{t1})-(\ref{t2}) where $\Theta$ is solution of an implicit scalar equation. \begin{theorem}\label{exist1} Under the assumptions of Theorem \ref{exist}, there exist weak solutions $(\theta_k,\phi_k)\in W^{2,pn/(p+n)}(\Omega)\times W^{1,p}(\Omega)$, with $p>n$, to the variational problem (\ref{heatj})-(\ref{pbphij}), such that \begin{eqnarray*} \theta_k\rightarrow \Theta \quad\mbox{ in } H^1(\Omega), \qquad \phi_k\rightharpoonup\phi \mbox{ in $H^1(\Omega)$,} \end{eqnarray*} with $\Theta$ solving the implicit scalar equation \begin{equation}\label{tt} \Theta\int_{\partial\Omega}\alpha(\cdot,\Theta)hds =\int_\Omega\sigma(\cdot,\Theta)|\nabla\phi|^2dx+\int_{\Omega}gdx \end{equation} and $\phi$ solving (\ref{t1})-(\ref{t2}). Moreover \begin{equation}\label{tt1} {\sigma_\#\int_\Omega |\nabla\phi|^2dx+\int_{\Omega}g dx \over\alpha^\#\int_{\partial\Omega}|h|ds} \leq \Theta\leq{\sigma^\#\int_\Omega |\nabla\phi|^2dx+\int_{\Omega}gdx \over \alpha_\#h_\#|\Gamma|+\alpha^\#h_1|\Sigma|} . \end{equation} \end{theorem} \begin{remark} In face of (\ref{cc}) if the Seebeck coefficient is only a function on the temperature then (\ref{tt}) reads $$ 0=\int_\Omega\sigma(\cdot,\Theta)|\nabla\phi|^2dx+\int_{\Omega}gdx.$$ The presence of a generic heat source invalids the limit model (\ref{t1})-(\ref{t2}) connected with the original thermoelectric problem introduced in \cite{zamm}. \end{remark} The above results can be proved if the convexity of $\Omega$ is replaced by weaker assumptions, for instance when $\Omega$ is a plane bounded domain with Lipschitz and piecewise $C^2$ boundary whose angles are all convex \cite[p. 151]{grisv}, or when $\Omega$ is a plane bounded domain with curvilinear polygonal $C^{1,1}$ boundary whose angles are all strictly convex \cite[p. 174]{grisv}. For general bounded domains $\Omega$ of $\mathbb R^n$ with Lipschitz boundary, it is known that the integrability exponents for the gradients of the potential and temperature solutions may be larger than 3 \cite{elsch,zan}, if the restriction to the case of uniformly continuous coefficients in (\ref{heatj})-(\ref{pbphij}) is assured. However, a generalization for such nonsmooth domains of Theorem \ref{exist} and its limit model is not a direct consequence. Indeed, new proofs will be needed because the compact embedding $W^{2,pn/(p+n)}(\Omega)\hookrightarrow\hookrightarrow W^{1,p}(\Omega)$ is crucial to provide the weak continuity of the operator in the fixed point argument. For the two-dimensional limit model, let us show the existence result under the minimal regularity on $\sigma$. \begin{theorem}\label{teo1} Assume $n= 2$ and (H1)-(H4) hold. Then the problem (\ref{pbu})-(\ref{pbn}) has a weak solution in the sense of Definition \ref{d1}, for some $p> 2=n$, under sufficiently small data. \end{theorem} \begin{proposition}\label{exist2} Under the assumptions of Theorem \ref{teo1}, there exists $\phi$ solving (\ref{t1})-(\ref{t2}) with $\Theta$ solving the implicit scalar equation (\ref{tt}). \end{proposition} The study of the existence of three-dimensional weak solutions to the variational problem (\ref{heatj})-(\ref{pbphij}), under the assumption that the mapping $x\in\Omega\mapsto \sigma(x,T)$ is discontinuous for every $T\in \mathbb{R}$, is still an open problem. When the coefficient $\sigma$ of the principal part of the divergence form elliptic equation (\ref{defphi}) is a discontinuous function on the spatial variable, it invalidates the smoothness of the solution as is carried out in the literature \cite{cfl,kp,lsw,mey,rag}. \section{Auxiliary results} \label{aux} The existence of a unique solution $\phi\in H^1(\Omega)$, such that $\int_{\partial\Omega} \phi ds=0$, to an auxiliary problem is consequence of Lax-Milgram lemma, for details see \begin{proposition}[{\cite[Theorem 4.1]{zamm}}]\label{twp2} Let the assumptions (H2) and (\ref{amm}) or (\ref{amm-}) be fulfilled. Assume that $n\geq 2$, $\Omega\in C^{0,1}$, $\xi\in H^1(\Omega)$ and $h\in L^{p}(\partial\Omega)$ verify (\ref{cc}) for $p> 2(n-1)/n$. Then there exists a unique weak solution $\phi\in H^1(\Omega)$, such that $\int_{\partial\Omega}\phi ds=0$, to the variational problem \begin{equation}\label{auxphi} \int_\Omega\sigma(\xi)\nabla\phi\cdot \nabla\eta dx=-\int_\Omega\sigma(\xi)\alpha(\xi)\nabla\xi\cdot\nabla\eta dx +\int_{\partial\Omega} h\eta ds, \end{equation} for all $\eta \in H^1(\Omega)$ and in particular for all $\eta\in W^{1,p'}(\Omega)$. \end{proposition} Next, we establish some regularity for the potential auxiliary solution. \begin{proposition}\label{twp1} Let $p>n$, $h\in L^p(\partial\Omega)$ verify (\ref{cc}), $\xi\in W^{2,q}(\Omega)$ with $q=pn/(p+n)$, (H4) be fulfilled, and $\phi\in H^{1}(\Omega)$ solve the problem (\ref{auxphi}). If $ \sigma:\Omega\times\mathbb R\rightarrow\mathbb R$ is a uniformly continuous function verifying (\ref{smin}), then $\nabla\phi\in {\bf L}^{p}(\Omega)$ and it verifies \begin{equation} \|\nabla\phi\|_{p,\Omega} \leq C(n,p,\Omega,\sigma_\#,\sigma^\#)(\sigma^\#\alpha^\#\|\nabla\xi\|_{p,\Omega}+\|h\|_{p,\partial\Omega} ) .\label{cotap} \end{equation} \end{proposition} {\sc Proof.} For $p>n$, we have $q>n/2$, \begin{eqnarray*} \xi\in W^{2,q}(\Omega)\hookrightarrow C^{0,2-n/q}(\bar\Omega),\quad \alpha(\xi) \nabla\xi\in {\bf W}^{1,q}(\Omega) \hookrightarrow{\bf L}^{q(n-1)/(n-q)}(\partial\Omega),\\ \alpha(\xi)\nabla\xi\cdot{\bf n}\in { L}^{q(n-1)/(n-q)}(\partial\Omega) \equiv { L}^{p(n-1)/n}(\partial\Omega)\hookrightarrow {W}^{-1/p,p}(\partial\Omega), \end{eqnarray*} for ${\bf n}\in{\bf L}^\infty(\partial\Omega)$. Moreover, $h/\sigma(\xi)\in W^{-1/p,p}(\partial\Omega)$ and $\nabla\xi\in {\bf L}^p(\Omega)$ implies that $\nabla\cdot(\sigma( \xi)\alpha(\xi)\nabla\xi)\in \left({\bf W}^{1,p'}(\Omega)\right)'$ for $p>n$. By appealing to the regularity theory \cite{daut1,grisv} for the solution $\phi\in H^{1}(\Omega)$ of the boundary value problem (in the sense of distributions) \begin{eqnarray*} &&\nabla\cdot\left(\sigma(\xi)\nabla \phi+{ \sigma(\xi)}\alpha(\xi)\nabla\xi\right)=0\quad \mbox{ in }\Omega\\ &&\sigma(\xi)(\nabla \phi+\alpha(\xi)\nabla\xi)\cdot{\bf n}=h\quad\mbox{on }\partial\Omega, \end{eqnarray*} and observing that $\xi\in W^{1,p}(\Omega)\hookrightarrow C^{0,1-n/p}(\bar\Omega)$ for $p>n$ warrants that $\sigma(\cdot,\xi)$ is uniformly continuous, then the regularity of weak solutions relative to $W^{1,p}(\Omega)$ and the estimate (\ref{cotap}) arise. $\Box$ The following result deals with the existence and uniqueness of a strong temperature auxiliary solution. \begin{proposition}\label{propt} Let $p>n$, $\xi\in W^{1,p}(\Omega)$, (H1) and (H3)-(H4) be fulfilled, and $\phi\in H^{1}(\Omega)$ solve the problem (\ref{auxphi}). If $\sigma:\Omega\times\mathbb R\rightarrow\mathbb R$ is a uniformly continuous function verifying (\ref{smin}), then there exists a unique weak solution $\theta\in W^{2,pn/(p+n)}(\Omega)$ solving the problem, for all $\eta \in W^{1,p'}(\Omega)$, \begin{eqnarray} k\int_\Omega \nabla\theta\cdot\nabla\eta dx +\int_{\partial\Omega}\alpha(\xi)h\theta \eta ds = \int_\Omega \left(F(\cdot,\xi,\nabla\xi,\nabla\phi) +g\right) \eta dx, \label{auxu}\end{eqnarray} with $F:\Omega\times\mathbb R^{2n+1}\rightarrow\mathbb R$ defined as $F(x,T,{\bf a},{\bf b})=$ \[\sigma(x,T)\Big(\alpha(x,T)(\alpha(x,T)+ {\partial\alpha\over\partial T}(x,T)T)|{\bf a}|^2+(2\alpha(x,T)+{\partial\alpha\over\partial T}(x,T)T){\bf a} \cdot{\bf b}+|{\bf b}|^2\Big).\] \end{proposition} {\sc Proof.} The existence and uniqueness of $\phi\in W^{1,p}(\Omega)$ is consequence of Propositions \ref{twp2} and \ref{twp1}. By appealing to the elliptic equations theory \cite{grisv}, from $F(\xi,\nabla\xi,\nabla\phi)+g\in L^ {p/2}(\Omega)$, the regularity theory for the Laplace equation in convex domains guarantees the existence of a unique solution $\theta\in W^{2,pn/(p+n)}(\Omega)$ of the Robin problem \begin{eqnarray*} &&-k\Delta \theta=F(\cdot,\xi,\nabla\xi,\nabla\phi)+g\quad \mbox{ in }\Omega;\\ &&k\nabla \theta\cdot{\bf n}+\alpha(\xi) h\theta=0\quad \mbox{ on }\partial\Omega, \end{eqnarray*} taking into account that the Korn perturbation method \cite[pp. 107-109]{grisv} can be adapted if the coefficient $\alpha(\cdot,\xi)h\in C(\partial \Omega)$ is such that the assumption (\ref{la}) holds. For this, we observe that $\xi\in W^{1,p}(\Omega)\hookrightarrow C(\bar\Omega)$ and we recall (H3)-(H4). $\Box$ For the regularity of the potential auxiliary solution $\phi$ when it is the unique weak solution for Neumann problem to an elliptic second order equation in divergence form with bounded and measurable coefficient, we can prove the following result. \begin{proposition}\label{pphi} If the assumptions of Proposition \ref{twp2} are fulfilled with $ p=2$ and $\Omega$ is convex, then there exists a constant $\epsilon>0$ such that the weak solution $\phi\in H^1(\Omega)$ of (\ref{auxphi}) belongs to $W^{1,2+\epsilon}(\Omega)$, i.e. \begin{equation}\label{cotaphie} \|\nabla {\phi}\|_{2+\epsilon,\Omega}\leq K_2(\sigma^\#\alpha^\#\|\nabla\xi\|_{2,\Omega}+\|h\|_{2,\partial\Omega} ) , \end{equation} with a constant $K_2>0$ only dependent on the data. \end{proposition} {\bf Proof.} Denote the operator $A$ by \[ \langle A\phi,\eta\rangle=\int_\Omega \sigma(\xi)\nabla\phi\cdot\nabla \eta dx. \] Then $\phi\in H^{1}(\Omega)$ is a weak solution to the second order elliptic differential equation $Au=F$, under \[F=\nabla\cdot( \sigma(\xi)\alpha(\xi)\nabla\xi)+h\in ( H^{1}(\Omega))'\hookrightarrow ( W^{1,p}(\Omega))',\qquad \forall p\geq 2.\] Since the boundedness property \[ \sigma_\#\leq \sigma(\cdot,\xi)\leq \sigma^\#, \mbox{ a.e. in }\Omega,\] is fulfilled, considering that $\xi\in L^1(\Omega)$ and the assumption (\ref{smin}) on $\sigma$ holds, then the Neumann version of the general result on the higher regularity for weak solutions to the mixed boundary value problems (cf. \cite[Theorem 1]{grog}, also \cite{grogr}) guarantees that $\phi\in W^{1,2+\epsilon}(\Omega)$ for some $\epsilon >0$. $\Box$ Although Proposition \ref{pphi} is valid for any dimensional space $(n\geq 2)$, we only used it for $n=2$. Let us precise its application in the following proposition. \begin{proposition}\label{propt2} Let $\xi\in W^{1,2+\epsilon}(\Omega)$, (H1)-(H4) be fulfilled, and $\phi\in H^{1}(\Omega)$ solve the problem (\ref{auxphi}). If the assumptions of Proposition \ref{pphi} hold, then there exists a unique weak solution $\theta\in W^{2,2p/(p+2)}(\Omega)$ solving (\ref{auxu}) with $p=2+\epsilon$. \end{proposition} {\sc Proof.} The imperative requirement of the embedding $ W^{1,2+\epsilon}(\Omega)\hookrightarrow C(\bar\Omega)$ yields that $\alpha(\cdot,\xi)h\in C(\partial \Omega)$ provided by (H3)-(H4). Thus, Proposition \ref{pphi} ensures that the argument of the proof of Proposition \ref{propt} is still valid, concluding the claim. $\Box$ \section{Proof of Theorem \ref{exist}} \label{sexist} First we recall the Tychonoff extension to weak topologies of the Schauder fixed point theorem \cite[pp. 453-456 and 470]{dsch}. \begin{theorem}\label{fpt} Let $K$ be a nonempty compact convex subset of a locally convex space $X$. Let ${\mathcal L}:K\rightarrow K$ be a continuous operator. Then $\mathcal L$ has at least one fixed point. \end{theorem} If we provide any Banach space with the weak topology, every closed ball is convex and weakly sequential compact. In order to apply Theorem \ref{fpt}, let us consider the operator $\mathcal L$ defined in a closed ball $\bar B_R \subset W^{2,pn/(p+n)}(\Omega)$ such that \[ \mathcal{L}:\xi\in\bar B_R\mapsto\phi\mapsto \theta\in W^{2,pn/(p+n)}(\Omega), \] where $\phi\in W^{1,p}(\Omega)$ solves the problem (\ref{auxphi}), for all $\eta \in W^{1,p'}(\Omega)$, and $\theta$ solves the problem (\ref{auxu}). The existence of a unique solution $\phi\in W^{1,p }(\Omega)$, such that $\int_{\partial\Omega} \phi ds=0$, to the problem (\ref{auxphi}) is consequence of Propositions \ref{twp2} and \ref{twp1}, and it verifies (\ref{cotap}). Hence, for $p>n$, we find $\theta\in W^{1,p}(\Omega)$ from Proposition \ref{propt}, and the estimate \begin{equation} \| \theta\|_{2,pn/(p+n),\Omega} \leq K(A\|\nabla\xi\|_{p,\Omega}^2+B\|\nabla\xi\|_{p,\Omega} \|\nabla\phi\|_{p,\Omega}+\sigma^\#\|\nabla\phi\|_{p,\Omega}^2+ \|g\|_{p/2,\Omega}), \label{cotatp} \end{equation} is verified with $K$ denoting a constant dependent on $\Omega$, $n$ and $p$, $A$ and $B$ given by (\ref{ab}), and $\varkappa:=\min\{k,\alpha_\#h_\#\}+\alpha^\# h_1C_1>0$. Thus $\mathcal L$ is well defined. Next, let us prove that $\mathcal L(\bar B_R) \subset \bar B_R.$ Let $\xi\in \bar B_R$ be arbitrary and $(\phi,\theta)$ be the corresponding solution solving (\ref{auxphi}) and (\ref{auxu}). Thus (\ref{cotap}) and (\ref{cotatp}) read \begin{eqnarray} \|\phi\|_{1,p,\Omega} \leq {C}(\sigma^\#\alpha^\#R+\|h\|_{p,\partial\Omega} ) ;\label{cotapr}\\ \label{cotatpr} \| \theta\|_{2,pn/(p+n),\Omega} \leq K(AR^2+BR \|\nabla\phi\|_{p,\Omega}+\sigma^\#\|\nabla\phi\|_{p,\Omega}^2+ \|g\|_{p/2,\Omega}), \end{eqnarray} with $C=C(n,p,\Omega,\sigma_\#,\sigma^\#)$. Inserting (\ref{cotapr}) into (\ref{cotatpr}) it follows \[\| \theta\|_{2,pn/(p+n),\Omega}\leq a_2R^2+a_1R+a_0, \] where \begin{eqnarray*} a_2&=& K\sigma^\#\alpha^\#\left(1+{C\sigma^\#}\right) \left(\alpha^\#\left(2+{C\sigma^\#}\right)+\mu^\#\right); \\a_1&=& {KC\sigma^\#}\left(2\alpha^\#(1+ \sigma^\#)+\mu^\#\right) \|h\|_{p,\partial\Omega}; \\a_0&=& K\left({C^2\sigma^\#}\|h\|_{p,\partial\Omega}^2+ \|g\|_{p/2,\Omega}\right). \end{eqnarray*} Therefore, $\mathcal L(\xi)=\theta\in\bar B_R$ if and only if $a_2R^2+(a_1-1)R+a_0\leq 0$, i.e. for instance if the smallness condition \begin{equation}\label{ss} a_1<1\quad\mbox{and}\quad 4a_0a_2<(1-a_1)^2 \end{equation} is assumed. \subsection{The weak sequential continuity of $\mathcal L$} \label{swc} Let $\{\xi_m\}_{m\in\mathbb N}$ be a sequence in $\bar B_R$ verifying \begin{equation}\label{xim2} \xi_m\rightharpoonup \xi\quad\mbox{in } W^{2,pn/(p+n)}(\Omega)\hookrightarrow\hookrightarrow W^{1,p}(\Omega), \end{equation} and $({\phi}_m,\theta_m)$ is the correspondent solution to (\ref{auxphi}) and (\ref{auxu}), for each $m\in \mathbb N$. From the estimates (\ref{cotap}) and (\ref{cotatp}) we can extract a subsequence, still labeled by $(\phi_m,\theta_m)$, such that \[ \phi_m \rightharpoonup \phi\quad \mbox{ in }W^{1,p}(\Omega),\qquad {\theta}_m \rightharpoonup{\theta}\quad \mbox{ in } W^{2,pn/(p+n)}(\Omega)\hookrightarrow\hookrightarrow W^{1,p}(\Omega). \] Thanks to Remark \ref{mpnl} it follows \begin{equation}\label{xim1} \xi_m \rightarrow \xi, \quad {\theta}_m \rightarrow{\theta},\quad \phi_m\rightarrow\phi\quad \mbox{ in }C^{0,1-n/p}(\bar\Omega). \end{equation} In particular, $\int_{\partial\Omega}\phi_m ds=0\rightarrow \int_{\partial\Omega} \phi ds=0$. By the continuity of the Nemytskii operators $\alpha$ and $\sigma$, we can pass to the limit in (\ref{auxphi})$_m$ as $m$ tends to infinity, concluding that $\phi\in W^{1,p}(\Omega)$ is the limit solution, i.e. it verifies (\ref{auxphi}). In the sequel, let us pass to the limit in (\ref{auxu})$_m$ as $m$ tends to infinity. First, the mapping $\xi\in L^1(\Omega)\mapsto \alpha(\xi)\in L^r(\Omega),$ for all $r <+\infty,$ is continuous by (H4), thus the passage to the limit to the left hand side of (\ref{auxu}) is straightforward. In order to study the RHS, we define \begin{eqnarray*} I_{1,m}&=& \int_\Omega \sigma(\xi_m)\alpha(\xi_m)(\alpha(\xi_m)+ {\partial\alpha\over\partial T}(\xi_m)\xi_m)|\nabla\xi_m|^2\eta dx;\\ I_{2,m}&=&\int_\Omega\sigma(\xi_m)(2\alpha(\xi_m)+{\partial\alpha\over\partial T}(\xi_m)\xi_m) \nabla\xi_m\cdot\nabla\phi_m\eta dx;\\ I_{3,m}&=& \int_\Omega {\sigma(\xi_m)}|\nabla\phi_m|^2\eta dx. \end{eqnarray*} Recalling Remark \ref{mpnl} we get $\eta\in W^{1,p'} (\Omega)\hookrightarrow L^{pn/(pn-n-p)}(\Omega)\hookrightarrow L^{p/(p-2)}(\Omega)$ for $p>n$. From (\ref{xim2}) we have $|\nabla{\xi}_m|^2 \rightarrow |\nabla\xi|^2$ in ${ L}^{p/2}(\Omega)$. Considering that the mapping $\xi\in L^1(\Omega)\mapsto \sigma(\xi)\alpha^2(\xi)\eta\in L^{p/(p-2) }(\Omega)$ is continuous thus the first term in $I_{1,m}$ passes to the limit as $m$ tends to infinity. Using (\ref{xim2}) and (\ref{xim1}) we have $\xi_m|\nabla{\xi}_m|^2 \rightarrow \xi|\nabla\xi|^2$ in ${ L}^{p/2}(\Omega)$. Considering that the mapping $\xi\in L^1(\Omega)\mapsto \sigma(\xi)\alpha(\xi){\partial\alpha\over\partial T}(\xi)\eta\in L^{p/(p-2) }(\Omega)$ is continuous thus the second term in $I_{1,m}$ passes to the limit as $m$ tends to infinity. Analogously, we take to the limit in $I_{2,m}$ observing that the strong-weak convergence product $\nabla{\xi}_m\cdot\nabla\phi_m\rightharpoonup \nabla\xi\cdot\nabla\phi$ holds in ${ L}^{p/2}(\Omega)$. In order to be in conditions for finding that $\theta$ is a limit solution, let us prove the continuity of the solution mapping $\xi\in W^{1,p}(\Omega)\mapsto\phi=\phi(\xi)\in W^{1,s}(\Omega)$ in the strong topology for $s=2pn/(p+n)<p$. Take the difference of (\ref{auxphi})$_m$ and (\ref{auxphi}) verified by the solutions $\phi_m$ and $\phi$, respectively, and choose $\eta=\phi_m-\phi$ as a test function. Thus, it results \begin{eqnarray*} \sigma_\#\|\nabla (\phi_m-\phi)\|_{2,\Omega}^2\leq \int_\Omega (\sigma(\xi)-\sigma(\xi_m))\nabla\phi\cdot\nabla(\phi_m-\phi) dx+\\ +\int_\Omega (\sigma(\xi)\alpha(\xi)\nabla\xi-\sigma(\xi_m)\alpha(\xi_m)\nabla\xi_m) \cdot\nabla(\phi_m-\phi) dx\longrightarrow 0, \quad\mbox{as $m\rightarrow\infty$.} \end{eqnarray*} Then, we conclude that $\nabla{\phi}_m \rightarrow\nabla\phi$ in ${\bf L}^{2}(\Omega)$, and consequently $\nabla{\phi}_m \rightarrow\nabla\phi$ a.e. in $\Omega$ and $|\nabla{\phi}_m|^2 \rightarrow |\nabla\phi|^2$ in ${L}^{s/2}(\Omega)\hookrightarrow L^{pn/(p+n)}(\Omega)$. Thus $I_{3,m}$ passes to the limit as $m$ tends to infinity, concluding the proof of weak continuity of the operator $\mathcal L$. Then the Schauder fixed point theorem can be used and it guarantees the existence of $(\theta,\phi)$ in the conditions to Theorem \ref{exist}. \subsection{The validation of the estimates (\ref{cotaphi})-(\ref{cotau})} Let $(\theta,\phi)\in W^{2,pn/(p+n)}(\Omega)\times W^{1,p}(\Omega)$ be a weak solution to the variational problem (\ref{heatj})-(\ref{pbphij}). Choose $\eta=\phi\in W^{1,p}(\Omega)$ as a test function in (\ref{pbphij}). Using (\ref{smin}), the upper bound of $|\alpha|$ and the Sobolev-Poincar\'e inequality then (\ref{cotaphi}) holds. From the regularity theory for the Robin-Laplace problem and by virtue of the existence of a solution $\theta\in W^{2,pn/(p+n)}(\Omega)$ we proceed as in (\ref{cotatp}) now for $k>\alpha_\#h_\#$ resulting the estimate \[ \| \nabla\theta\|_{p,\Omega} \leq K(A\|\nabla\theta\|_{p,\Omega}^2+B\|\nabla\theta\|_{p,\Omega} \|\nabla\phi\|_{p,\Omega}+\sigma^\#\|\nabla\phi\|_{p,\Omega}^2+ \|g\|_{p/2,\Omega}), \] with $K$ denoting a constant independent on $k$. Combining this result with the estimate (\ref{cotap}) with $\xi$ replaced by $\theta$ and using (\ref{ss}) we conclude (\ref{cotaup}). Choose $\eta=\theta\in W^{1,p}(\Omega)$ as a test function in (\ref{heatj}). Then applying the H\"older inequality and using the assumptions (H1)-(H4) it follows \begin{eqnarray*} k\|\nabla \theta\|^2_{2,\Omega}\leq\alpha^\#h_1 \|\theta\|^{2}_{2,\Sigma}+ \Big(A\|\nabla\theta\|_{2,\Omega}^2+B\|\nabla \theta\|_{2,\Omega}\|\nabla\phi\|_{2,\Omega}+\\+\sigma^\#\|\nabla \phi\|_{2,\Omega}^2+ \|g\|_{1,\Omega}\Big)\|\theta\|_{\infty,\Omega}. \end{eqnarray*} This yields the estimate (\ref{cotau}). \section{Proof of Theorem \ref{exist1}} \label{sexist1} For each given $k>0,$ let $(\theta_k,\phi_k)$ be a solution to (\ref{heatj})-(\ref{pbphij}) in accordance with Theorem \ref{exist}. From estimates (\ref{cotau}) and (\ref{cotaphi}) there exist subsequences still denoted by $\theta_k$ and $\phi_k$ such that, for $k\rightarrow+\infty,$ \begin{eqnarray*} \nabla \theta_k\rightarrow 0 \quad\mbox{ in }{\bf L}^2(\Omega);\qquad \theta_k\rightharpoonup \Theta \quad\mbox{ in } W^{1,p}(\Omega)\hookrightarrow\hookrightarrow C(\bar\Omega);\\ \phi_k\rightharpoonup\phi\qquad \mbox{ in $H^1(\Omega)$}, \end{eqnarray*} with $\Theta$ constant on $\bar\Omega$. Hence we can pass to the limit in (\ref{pbphij}) as $k$ tends to infinity resulting \[ \int_\Omega\sigma(\cdot,\Theta)\nabla\phi\cdot \nabla\eta dx= \int_{\partial\Omega} h\eta ds,\quad\forall \eta\in W^{1,p'}(\Omega), \] or equivalently (\ref{t1})-(\ref{t2}). In particular, if we take $\eta=1$ in (\ref{heatj}) we can pass to the limit as $k$ tends to infinity resulting (\ref{tt}). Using (\ref{smin}) it follows \begin{eqnarray*} \sigma_\#\int_\Omega|\nabla\phi|^2dx+\int_\Omega gdx \leq \Theta\int_{\partial\Omega}\alpha(\cdot,\Theta)h ds\leq \sigma^\#\int_\Omega|\nabla\phi|^2dx+\int_\Omega gdx. \end{eqnarray*} Taking into account that the assumption (H4)$_+$ or (H4)$_-$ and also (\ref{c1}) imply \begin{eqnarray*} \int_{\partial\Omega}\alpha(\cdot,\Theta)h ds &\leq &\alpha^\# \int_{\partial\Omega}|h|ds;\\ \int_{\partial\Omega}\alpha(\cdot,\Theta)h ds &\geq & \alpha_\#h_\#|\Gamma|+\alpha^\# h_1|\Sigma|>0, \end{eqnarray*} we derive (\ref{tt1}), concluding the proof of Theorem \ref{exist1}. \section{The two-dimensional limit model} \label{2d} \subsection{Proof of Theorem \ref{teo1}} Arguing as in Theorem \ref{exist}, the Schauder fixed point argument can be applied. For this, it is sufficient to see that the regularity relative to $W^{1,p}_{\rm loc}(\bar\Omega)=W^{1,p}(\Omega)$ ($\Omega$ bounded) can be applied for the unique weak solution of the variational problem (\ref{auxphi}) for $p=2+\epsilon>2=n$ in accordance to Proposition \ref{pphi}. Thus Proposition \ref{propt2} guarantees the existence of $\theta\in W^{2, 2p/(p+2)}(\Omega)$ verifying (\ref{cotatp}). For every $\xi\in \bar B_R$, inserting (\ref{cotaphie}) into (\ref{cotatp}) it follows \begin{eqnarray*} \| \theta\|_{2,2p/(p+2),\Omega} \leq K(AR^2+BR {K_2}(\sigma^\#\alpha^\#R+\|h\|_{2,\partial\Omega} ) +\\ +\sigma^\# {K_2}^2(\sigma^\#\alpha^\#R+\|h\|_{2,\partial\Omega} ) ^2+ \|g\|_{p/2,\Omega}). \end{eqnarray*} Next arguing as in Section \ref{sexist} it leads to a smallness condition. From the continuity of the Nemytskii operator $\sigma $ due to the Krasnoselski Theorem we can proceed as in Section \ref{swc}, considering that $I_{3,m}$ passes to the limit as $m$ tends to infinity since \begin{eqnarray*} \sigma(\xi_m)\rightharpoonup \sigma(\xi)\qquad \mbox{weakly* in }L^\infty(\Omega);\\ |\nabla{\phi}_m|^2\eta \rightarrow |\nabla\phi|^2\eta\qquad \mbox{in }L^{1}(\Omega). \end{eqnarray*} This concludes that we are in the conditions of applying Theorem \ref{fpt}. \subsection{Proof of Proposition \ref{exist2}} We can proceed as in Section \ref{sexist1} considering the existence of the sequence of solutions is provided by Theorem \ref{teo1} and the proposition follows.
\section{Statement of Main Result} In a previous paper \cite{piwkb}, the author studied the distribution of poles of solutions of the the first Painlev\'e equation \begin{equation*} y''= 6y^2 -z \, , \; z \in \mathbb{C} \quad , \end{equation*} with a particular attention to the poles of the int\'egrale tritronqu\'ee. This is the unique solution of P-I with the following asymptotic behaviour at infinity \begin{equation*} y(z) \sim - \sqrt{\frac{z}{6}}, \quad \mbox{if} \quad |\arg z| <\frac{4 \pi}{5} \; . \end{equation*} The problem of computing the poles of the tritronqu\'ee solution was mapped to a pair of spectral problems for the cubic anharmonic oscillator. More precisely, it was shown that a point $a \in \mathbb{C}$ is a pole of the tritronqu\'ee solution if and only exists $b \in \mathbb{C}$ such that the following Schr\"odinger equation \begin{equation}\label{eqn:schroedinger} \frac{d^2\psi(\lambda)}{d\lambda^2}= V(\lambda;a,b) \psi(\lambda)\; , \quad V(\lambda;a,b)=4 \lambda^3 -2 a \lambda -28 b \;. \end{equation} admits the simultaneous solutions of two different quantization conditions. Using a suitable complex WKB method, the author studied this pair of quantization conditions. He derived a system of two equations, the Bohr-Sommerfeld-Boutroux (B-S-B) system, whose solutions describe approximately the distribution of the poles. \begin{figure}[htbp] \begin{center} \input{figure1.pstex_t} \end{center} \caption{Riemann surface $\mu^2=V(\lambda;a,b)$} \label{figure:riemann} \end{figure} We say that $(a,b)$ satisfies the Bohr-Sommerfeld-Boutroux (for the precise definition see Definition \ref{def:bsb} below) system if \begin{eqnarray}\nonumber \oint_{c_{- 1}}\sqrt{V(\lambda;a,b)}d\lambda &=& i\pi (2n-1) \; , \\ \label{sys:bsb} \\ \nonumber \oint_{c_{1}}\sqrt{V(\lambda;a,b)}d\lambda &=& i\pi (2m-1)\; . \end{eqnarray} Here $m,n \in \bb{N}-0$ are called quantum numbers and the cycles $c_{\pm1}$ are depicted in Figure \ref{figure:riemann}. For any pair of quantum numbers there is one and only one solution to the Bohr-Sommerfeld-Boutroux system; this is proven for example in \cite{kapaev03}. Solutions of B-S-B system have naturally a multiplicative structure. \begin{Def}\label{def:primitive} Let $(a^*, b^*)$ be a solution of the B-S-B system with quantum numbers $n,m$ such that $2n-1$ and $2m -1$ are coprime. We call $(a^*,b^*)$ a {\rm primitive} solution of the system and denote it $(a^q,b^q)$, where $q=\frac{2n-1}{2m-1} \in \bb{Q}$. Due to Lemma \ref{lem:chiderivatives} below, we have that $$(a_k^q,b_k^q)=((2k +1)^{\frac{4}{5}}a^q,(2k+1)^{\frac{6}{5}}b^q) , k \in \bb{N} \; ,$$ is another solution of the B-S-B system. We call it a {\rm descendant} solution. We call $\left\lbrace (a_k^q,b_k^q) \right\rbrace_{k \in \bb{N}}$ the q-sequence of solutions. \end{Def} In \cite{piwkb} it is shown that the sequence of real solutions of B-S-B system is the 1-sequence of solutions. The real primitive solution is computed numerically as $a^1 \cong -2,34, b^1 \cong -0,064 $. In the present paper we prove that any q-sequence approximates a sequence of poles of the tritronqu\'ee solution. The error between the pole and its WKB estimate is of order $(2k+1)^{-\frac{6}{5}}$ (see Theorem \ref{thm:aapprox} below). \begin{Def}\label{def:disc} We denote $D_{\epsilon}(a')=\left\lbrace\av{a-a'} <\varepsilon , \, \varepsilon \neq 0\right\rbrace$. \end{Def} The main results of the present paper is the following \begin{Thm}[Main Theorem]\label{thm:aapprox} Let $\varepsilon$ be an arbitrary positive number. If $\frac{1}{5} <\alpha < \frac{6}{5}$, then it exists $K \in \bb{N}^*$ such that for any $k\geq K$ inside the disc $D_{k^{-\alpha}\varepsilon}(a^q_k)$ there is one and only one pole of the int\'egrale tritronqu\'ee. \end{Thm} The rest of the paper is devoted to the proof of the theorem. \section{Proof} \subsection{Multidimensional Rouche Theorem} The main technical tool of the proof is the following generalization of the classical Rouch\'e theorem. \begin{Thm}[\cite{aizenberg}] Let $D,E$ be bounded domains in $\bb{C}^n $, $ \overline{D} \subset E$ , and let $f(z) , g(z)$ be holomorphic maps $ E \to \bb{C}^n$ such that \begin{itemize} \item $f(z) \neq 0 ,\; \forall z \in \partial D$, \item $\av{g(z)} < \av{f(z)}, \;\forall z \in \partial D$, \end{itemize} then $w(z)=f(z)+g(z)$ and $f(z)$ have the same number (counted with multiplicities) of zeroes inside $D$. Here $\av{f(z)}$ is any norm on $\bb{C}^n$. \end{Thm} \subsection{Monodromy of Schr\"odinger Equation} Poles of int\'egrale tritronque\'ee are in bijections with the simultaneous solutions of two eigenvalues problems for the cubic anharmonic oscillator \cite{piwkb}. Below we recall the basics of anharmonic oscillators theory; all the details can be found in \cite{piwkb}. Fix $k \in \mathbb{Z}_5 = \left\lbrace -2, \dots,2 \right\rbrace$ and the branch of $\lambda^{\frac{1}{2}}$ in such a way that ${\rm Re}{\lambda^{\frac{5}{2}}} \to + \infty$ as $|\lambda| \to \infty,\arg{\lambda}=\frac{2\pi k}{5}$. Then there exists a unique solution $\psi_k(\lambda)$ of equation (\ref{eqn:schroedinger}) such that \begin{equation}\label{def:psik} \lim_{\lambda \to \infty, \av{\lambda -\frac{2 \pi k}{5}} < \frac{3 \pi }{5} -\varepsilon} \lambda^{\frac{3}{4}}e^{+\frac{4}{5}\lambda^{\frac{5}{2}} - \frac{1}{2} a\lambda^{\frac{1}{2}}}\psi_k(\lambda;a,b)=1 \; . \end{equation} For any pair of functions $\psi_l, \psi_{l+2}$, we call \begin{equation}\label{def:wk} w_k(l,l+2)=\lim_{\substack{\lambda\ \to \infty \\ \left|\arg{\lambda}-\frac{2 \pi k}{5}\right| < \frac{\pi}{5} -\varepsilon}}\frac{\psi_l(\lambda)}{\psi_{l+2}(\lambda)} \in \mathbb{C} \cup \infty \, , \; k \in \mathbb{Z}_5 \; . \end{equation} the k-th asymptotic value. If $\psi_{l}$ and $\psi_{l+2}$ are linearly independent then $w_k(l,l+2)=w_m(l,l+2)$ if and only if $\psi_k$ and $\psi_m$ are linearly dependent. \begin{Def} Let $E$ be the (open) subset of the $(a,b)$ plane such that $\psi_0(\lambda;a,b)$ and $\psi_{\pm2}(\lambda;a,b)$ are linearly independent (its complement in the $(a,b)$ plane is the union of two smooth surfaces \cite{eremenko}). On $E$ we define the following functions \begin{eqnarray} \label{def:u2} u_2(a,b) &=& \frac{w_{2}(0,-2)}{w_{-1}(0,-2)} \\ \label{def:u-2} u_2(a,b) &=& \frac{w_{-2}(0,2)}{w_{1}(0,2)} \\ \label{def:U} U(a,b) &=& \left( \begin{matrix} u_2(a,b)-1 \\ u_{-2}(a,b) -1 \end{matrix} \right) \; . \end{eqnarray} All the functions are well defined and holomorphic. Indeed, due to WKB theory we have that $w_{l+1}(l,l+2)$ is always different from $0$ and $\infty$. \end{Def} We can characterize the poles of the int\'egrale tritronqu\'e as the zeroes of $U$. \begin{Thm}[\cite{piwkb}]\label{lem:Uzero} The point $a \in \bb{C}$ is a pole of the int\'egrale tritronqu\'ee if and only if there exists $b \in \bb{C}$ such that $(a,b)$ belongs to the domain of $U$ and $U(a,b)=0$. In other words $\psi_{-1}(\lambda;a,b)$ and $\psi_2{(\lambda;a,b)}$ are linearly dependent and $\psi_{1}(\lambda;a,b)$ and $\psi_{-2}{(\lambda;a,b)}$ are linearly dependent. \end{Thm} We remember that the complex number $b$ in previous lemma is the coefficient of the quartic term in the Laurent expansion of the tritronqu\'ee solution around $a$ (see Section 2.2 in \cite{piwkb}). \subsection{WKB Theory} Let $V(\lambda;a,b)$ be the potential of equation (\ref{eqn:schroedinger}). We call turning point any zero of $V$. A Stokes line is any curve in the complex $\lambda$ plane along which the real part of the action is constant, such that at least one turning point belong to its boundary. The union of all the Stokes line and all turning points is called the Stokes complex of the potential. A Stokes complex is naturally a graph embedded in the complex plane. The Stokes graphs has been classified topologically in \cite{piwkb} and the graph of type "320" (see Figure \ref{figure:320}) was shown to be crucial to the approximate description of the poles of the int\'egrale tritronqu\'ee. \begin{figure}[htbp] \begin{center} \input{figure2.pstex_t} \end{center} \caption{Graph "320": dots on the circle represents asymptotic directions in the complex plane} \label{figure:320} \end{figure} \begin{Def} Let $(a^*,b^*)$ be a point such that the Stokes graph of $V(^.;a,b)$ is of type "320". On a sufficiently small neighborhood of $(a^*,b^*)$ we define the following analytic functions \begin{eqnarray}\label{def:chi} \chi_{\pm2}(a,b) &=& \oint_{c_{\mp 1}}\sqrt{V(\lambda;a,b)}d\lambda \; , \\ \label{def:tu} \tilde{u}_{\pm2}(a,b) &=& -e^{\chi_{\pm2}(a,b)} \; ,\\ \label{def:tU} \tilde{U}(a,b) &=& \left( \begin{matrix} \tilde{u}_2(a,b)-1 \\ \tilde{u}_{-2}(a,b) -1 \end{matrix} \right) \; . \end{eqnarray} The cycles $c_{\pm1}$ are depicted in Figure 1 and the branch of $\sqrt{V}$ is chosen such that $\mbox{Re}\sqrt{V(\lambda)} \to + \infty$ as $\lambda \to \infty$ along the positive semi-axis in the cut plane. \end{Def} \begin{Def}\label{def:bsb} We say that $(a,b)$ satisfies the Bohr-Sommerfeld-Boutroux (B-S-B) system if the Stokes graph of $V(^.;a,b)$ is of type "320" and \begin{eqnarray}\nonumber \chi_{2}(a,b) &=& \oint_{c_{- 1}}\sqrt{V(\lambda;a,b)}d\lambda = i\pi (2n-1) \; , \\ \nonumber & &~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~(\ref{sys:bsb}) \\ \nonumber \chi_{-2}(a,b) &=& \oint_{c_{1}}\sqrt{V(\lambda;a,b)}d\lambda = i\pi (2m-1) \; . \end{eqnarray} Here $m,n \in \bb{N}-0$ are called quantum numbers. Equivalently the B-S-B system can be written as $\tilde{U}(a,b)=0$. \end{Def} In \cite{kapaev03} the following lemma was proven. \begin{Lem}\label{lem:uniq} For any pair of quantum numbers $n,m \in \bb{N}-0$ there exists one and only one solution of the B-S-B system. \end{Lem} After the results of \cite{piwkb} Section 4.3, we can compare the functions $U$ and $\tilde{U}$ defined above. \begin{Lem}\label{lem:uapprox} Let $(a,b)$ be such that the Stokes graph is of type "320". There exists a neighborhood of $(a,b)$ and two continuous positive functions $\rho_{\pm2}$ such that $\chi_{\pm2}$ are holomorphic and \begin{equation}\label{app:basic} \av{\tilde{u}_{\pm2} - u_{\pm2}} \leq \frac{1}{2}(e^{2 \rho_{\pm2}}-1) \; . \end{equation} Moreover if $\rho_{\pm2} < \frac{\ln 3}{2}$ then $\psi_0$ and $\psi_{\pm2}$ are linearly independent. \end{Lem} We remark that in \cite{piwkb} $\rho_{\pm2}$ were denoted $\rho^0_{\pm2}$. Using classical relations of the theory of elliptic functions we have the following \begin{Lem}\label{lem:invertibility} The map $\tilde{U}$ defined in (\ref{def:tU}) is always locally invertible (hence its zeroes are always simple) and \begin{eqnarray*} \frac{\partial \chi_{2} }{\partial a}(a,b) \frac{\partial \chi_{-2} }{\partial b}(a,b) - \frac{\partial \chi_{-2} }{\partial a}(a,b) \frac{\partial \chi_{2} }{\partial b}(a,b)= -28 \pi i \;. \end{eqnarray*} \begin{proof} On the compactified elliptic curve $\mu^2=V(\lambda;a,b)$, consider the differentials $\omega_a=-\frac{\lambda d\lambda}{\mu}$ and $\omega_b=-\frac{d\lambda}{\mu}$. It is easily seen that $$\frac{\partial \chi_{\pm2} }{\partial a}(a,b)=\oint_{c_{\mp1}}\omega_a \, , \; \frac{\partial \chi_{\pm2} }{\partial b}(a,b)=14\oint_{c_{\mp1}}\omega_b \; .$$ Moreover we have that $$J\tilde{U}=\left( \frac{\partial \chi_{2} }{\partial a}(a,b) \frac{\partial \chi_{-2} }{\partial b}(a,b) - \frac{\partial \chi_{-2} }{\partial a}(a,b) \frac{\partial \chi_{2} }{\partial b}(a,b)\right)\tilde{u}_{2}\tilde{u}_{-2} \; ,$$ where $J\tilde{U}$ is the Jacobian of the map $\tilde{U}$. The statement of the lemma follows from the classical Legendre relation between complete elliptic periods of the first and second kind \cite{bateman2}. \end{proof} \end{Lem} Our aim is to locate the zeroes of $U$ (the poles of the int\'egrale tritronqu\'ee after Theorem \ref{lem:Uzero}) knowing the location of zeroes of $\tilde{U}$(the solutions of the B-S-B system). We want to find a neighborhood of a given solution of the B-S-B system inside which there is one and only one zero of $U$. Due to estimate (\ref{app:basic}) and Rouch\'e theorem, it is sufficient to find a domain on whose boundary the following inequality holds \begin{equation}\label{inequality:rouche} \frac{1}{2} \left( e^{2\rho_{2}}-1\right)\av{u_2} + \frac{1}{2} \left( e^{2\rho_{-2}}-1\right) \av{u_{-2}} < \av{1-\tilde{u}_2}+ \av{1-\tilde{u}_{-2}} \; . \end{equation} \subsubsection{Scaling Law} In order to analyze the important inequality (\ref{inequality:rouche}), we take advantage of the following scaling behaviour that was proven in \cite{piwkb} Section 4.4. \begin{Lem}\label{lem:chiderivatives} Let $(a^*,b^*)$ be such that the Stokes graph is of type "320" and $E$ be a neighborhood of $(a^*,b^*)$ such that the estimates (\ref{app:basic}) are satisfied. Then, for any real positive $x$ the point $(x^2a^*, x^3b^*)$ is such that the Stokes graph is of type "320" and in the neighborhood $E_X=\left\lbrace (x^2a,x^3b): (a,b) \in U \right\rbrace $ the estimates (\ref{app:basic}) are satisfied. Moreover for any $(a,b) \in E$ the following scaling laws are valid \begin{itemize} \item $\chi_{\pm2}(x^2a,x^3b)= x^{\frac{5}{2}}\chi_{\pm2}(a,b)$. \item $\frac{\partial^{(n+m)} \chi_{\pm2} }{\partial a^n \partial b^m}(x^2a,x^3b)=x^{\frac{5-4 n -6 m}{5}} \frac{\partial^{(n+m)} \chi_{\pm2} }{\partial a^n \partial b^m}(a,b)$. \item $\rho_{\pm2}(x^2a,x^3b)= x^{-\frac{5}{2}}\rho_{\pm2}(a,b)$. \end{itemize} \end{Lem} \section{Proof of the main theorem} From Lemma \ref{lem:chiderivatives} we can extract the leading behaviour of $\tilde{U}$ around real solutions of the B-S-B system. \begin{Lem}\label{lem:linearization} Let $z=(2k+1)^{\alpha}(a-a^q_k)$, $c_{\pm2}=\frac{\partial \chi_{\pm2} }{\partial a}(a^q,b^q)$, $w=(2k+1)^{\beta}(b-b^q_k)$, and $d_{\pm2}=\frac{\partial \chi_{\pm2} }{\partial b}(a^q,b^q)$. If $\alpha > \frac{1}{5}$ and $\beta > -\frac{1}{5}$, then \begin{eqnarray}\nonumber \tilde{u}_{2}(z,w)\! &=& \!1+ c_2 (2k+1)^{\frac{1}{5} -\alpha}z + d_2 (2k+1)^{-\frac{1}{5} -\beta}w + O((2k+1)^{-\gamma'}) \; , \\ \label{exp:us} \\ \nonumber \tilde{u}_{-2}(z,w) \! &=&\! 1+ c_{-2} (2k+1)^{\frac{1}{5} -\alpha}z + d_{-2} (2k+1)^{-\frac{1}{5} -\beta}w + O((2k+1)^{-\gamma'}) \; , \\ \nonumber &&\gamma' > -\frac{1}{5}+\alpha, \gamma'>\frac{1}{5} +\beta \; . \end{eqnarray} \begin{proof} It follows from Lemma \ref{lem:chiderivatives}. \end{proof} \end{Lem} \begin{Def}\label{def:polydisc} We denote $D_{\epsilon, \delta}^{(a',b')}=\left\lbrace\av{a-a'} <\varepsilon , \av{b-b'}<\delta, \, \varepsilon,\delta \neq 0\right\rbrace$ the polydisc centered at $(a',b')$. \end{Def} We have collected all the elements for proving the following \begin{Lem}\label{lem:polydisc} Let $\varepsilon$, $\delta$ be arbitrary positive numbers. If $\frac{1}{5} <\alpha < \frac{6}{5}$, $-\frac{1}{5} <\beta < \frac{4}{5}$, then there exists a $K \in \bb{N}^*$ such that for any $k\geq K$, $U$ and $\tilde{U}$ are well-defined and holomorphic on $D_{k^{-\alpha}\varepsilon,k^{-\beta} \delta}^{(a^q_k,b^q_k)}$ and the following inequality holds true \begin{equation}\label{ineq:UU} \av{U(a,b) - \tilde{U}(a,b)} < \av{\tilde{U}(a,b)}, \forall (a,b) \in \partial D_{k^{-\alpha}\varepsilon,k^{-\beta} \delta}^{(a_k,b_k)} \; . \end{equation} \begin{proof} The polydisc $D_{k^{-\alpha}\varepsilon,k^{-\beta} \delta}^{(a^q_k,b^q_k)}$ is the image under rescaling $a \to (2k+1)^{\frac{4}{5}}a$, $b \to (2 k+1)^{\frac{6}{5}}b $ of a shrinking polydisc centered at $(a^q,b^q)$; call it $\tilde{D}_k$. Hence due to Lemma \ref{lem:uapprox}, for $k\geq K'$ $\tilde{D}_k$ is such that $\rho_{\pm2}$ are bounded, $\chi{\pm2}$ are holomorphic and the estimates (\ref{app:basic}) hold. Call $\rho^*$ the supremum of $\rho_{\pm2}$ on $D_{K'}$. Due to scaling property, for all $k \geq K'$ $\rho_{\pm2}$ is bounded from above by $(2k+1)^{-1}\rho^*$ on $D_{k^{-\alpha}\varepsilon,k^{-\beta} \delta}^{(a_k,b_k)}$; such a bound is eventually smaller than $\frac{\ln 3}{2}$. Then for a sufficiently large $k$, $D_{k^{-\alpha}\varepsilon,k^{-\beta} \delta}^{(a_k,b_k)}$ is a subset of the domain of $U$ and inside it $U$ and $\tilde{U}$ satisfy (\ref{app:basic}) and (\ref{exp:us}). We divide the boundary in two subsets $\partial D_{k^{-\alpha}\varepsilon,k^{-\beta}\delta}^{(a^q_k,b^q_k)}=D_0 \cup D_1$, $$D_0=\left\lbrace |a-a^q_k|=k^{-\alpha}\varepsilon; \av{b-b^q_k} \leq k^{-\beta}\delta\right\rbrace \,, $$ $$ D_1 = \left\lbrace |a-a_k| \leq k^{-\alpha}\varepsilon; \av{b-b_k} = k^{-\beta}\delta\right\rbrace \; .$$ Inequality (\ref{ineq:UU}) will be analyzed separately on $D_0$ and $D_1$. If $\av{d_2}\leq \av{d_{-2}}$, denote $d_2=d, d_{-2}=D, c=c_2, C=c_{-2}$; in the opposite case $\av{d_2} > \av{d_{-2}}$, denote $d_{-2}=d, d_{2}=D, c=c_{-2}, C=c_{2}$. By the triangle inequality and expansion (\ref{exp:us}), we have that $$ \av{\tilde{U}(a,b)} \geq (2k +1)^{\frac{1}{5} - \alpha} \varepsilon \av{(c - \frac{C d}{D})} + \mbox{ higher order terms } , \; (a,b) \in D_0 . $$ Similarly, if $\av{c_2}\leq \av{c_{-2}}$ denote $d_2=d, d_{-2}=D, c=c_2, C=c_{-2}$; in the opposite case $\av{c_2} > \av{c_{-2}}$, denote $d_{-2}=d, d_{2}=D, c=c_{-2}, C=c_{2}$. By the triangle inequality and expansion (\ref{exp:us}), we have that $$ \av{\tilde{U}(a,b)} \geq (2k +1)^{-\frac{1}{5} - \beta}\delta \av{(d - \frac{D c}{C})} + \mbox{ higher order terms }, \; (a,b) \in D_1 . $$ We observe that $(c - \frac{C d}{D})\neq0$ and $ (d - \frac{D c}{C})\neq0$, since (see Lemma \ref{lem:invertibility}) $c_2d_{-2}- c_{-2}d_2 = - 28 \pi i $. By hypothesis $ -1<\frac{1}{5} - \alpha<0$ and $ -1<-\frac{1}{5} - \beta<0$. Conversely, $\av{U(a,b) - \tilde{U}(a,b)} \leq \frac{\rho^*}{2k+1} + \mbox{ higher order terms},$ for all $(a,b) \in D_0\cup D_1$. The Lemma is proven. \end{proof} \end{Lem} As a corollary of Lemma \ref{lem:polydisc} and of Rouch\'e theorem, we obtain the following theorem which implies Theorem \ref{thm:aapprox}. \begin{Thm}\label{thm:approximation} Let $\varepsilon$, $\delta$ be arbitrary positive numbers. If $\frac{1}{5} <\alpha < \frac{6}{5}$, $-\frac{1}{5} <\beta < \frac{4}{5}$, then it exists a $K \in \bb{N}^*$ such that for any $k\geq K$, inside the polydisc $D_{k^{-\alpha}\varepsilon,k^{-\beta} \delta}^{(a^q_k,b^q_k)}$ there is one and only one solution of the system $U(a,b)=0$. \end{Thm} \paragraph{Acknowledgments} I am indebted to my advisor Prof. B. Dubrovin who constantly gave me suggestions and advice. \bibliographystyle{alpha}
\subsubsection{#1}} \newcommand{\subsectionlnk}[1]{\hypertarget{#1_subsec}{}\subsection{#1}} \newcommand{\sectionlnk}[1]{\hypertarget{#1_sec}{}\section{#1}} \newcommand{\paragraphlnk}[1]{\hypertarget{#1_par}{}\paragraph{#1}} \newcommand{\chargec}[1]{({#1})^c} \newcommand{{\, .}}{{\, .}} \newcommand{{\, ,}}{{\, ,}} \newcommand{\, .}{\, .} \newcommand{{\text{sphaleron}}}{{\text{sphaleron}}} \newcommand{\ampabs}[2]{\abs{\mathcal{M}_0}_{{#1}\rightarrow{#2}}} \newcommand{\amp}[2]{{\mathcal{M}}_{{#1}\rightarrow{#2}}} \newcommand{{m_{\text{Pl}}}}{{m_{\text{Pl}}}} \newcommand{\scalefac}{R} \newcommand{\speciesdist}[2]{f^{#1}_{#2}} \newcommand{\EQspeciesdist}[2]{{f^{#1,\,eq}_{#2}}} \newcommand{\speciesen}[2]{E^{#1}_{#2}} \newcommand{\deltafour}[1]{\delta (#1)} \newcommand{\lorentzd}[2]{\,d\Pi^{#1}_{#2}} \newcommand{\mybox}[1]{\mbox{\rule[-4mm]{0mm}{10mm}#1}} \newcommand{\partdiff}[3]{\left(\frac{\partial #1}{\partial #2}\right)_{#3}} \newcommand{\partd}[2]{\frac{\partial #1}{\partial #2}} \newcommand{\unit}[1]{\,\mbox{#1}} \newcommand{\order}[1]{\mathcal{O}(#1)} \newcommand{\mathcal{M}}{\mathcal{M}} \newcommand{\mathscr{L}}{\mathscr{L}} \newcommand{\,d}{\,d} \newcommand{\bvec{k}}{\bvec{k}} \newcommand{\bvec{p}}{\bvec{p}} \newcommand{\bvec{q}}{\bvec{q}} \newcommand{\bvec{r}}{\bvec{r}} \newcommand{\bvec{\lambda}}{\bvec{\lambda}} \newcommand{\bar{b}}{\bar{b}} \newcommand{\phi}{\phi} \newcommand{{q}}{{q}} \newcommand{{\ell}}{{\ell}} \newcommand{{q_L}}{{q_L}} \newcommand{{\ell_L}}{{\ell_L}} \newcommand{{q_R}}{{q_R}} \newcommand{{\ell_R}}{{\ell_R}} \newcommand{{q_{uR}}}{{q_{uR}}} \newcommand{{q_{uL}}}{{q_{uL}}} \newcommand{{q_{dR}}}{{q_{dR}}} \newcommand{{q_{uL}}}{{q_{uL}}} \newcommand{{\nu}}{{\nu}} \newcommand{{\nu_L}}{{\nu_L}} \newcommand{{N_R}}{{N_R}} \newcommand{{N_R}}{{N_R}} \newcommand{{N}}{{N}} \newcommand{{e}}{{e}} \newcommand{{e_R}}{{e_R}} \newcommand{{h.c.}}{{h.c.}} \newcommand{{SM\;}}{{SM\;}} \newcommand{\tr}{\mbox{Tr}} \providecommand{{m_{\text{Pl}}}}{{m_{\text{Pl}}}} \providecommand{\scalefac}{a} \providecommand{\hubblerate}{H} \providecommand{\speciesdist}[2]{f^{#1}_{#2}} \providecommand{\EQspeciesdist}[2]{{f^{#1}_{#2}}^{EQ}} \providecommand{\speciesen}[2]{E^{#1}_{#2}} \providecommand{\deltafour}[1]{\delta (#1)} \providecommand{\lorentzd}[2]{\,d\Pi^{#1}_{#2}} \providecommand{\mybox}[1]{\mbox{\rule[-4mm]{0mm}{10mm}#1}} \providecommand{{\text{sphaleron}}}{{\text{sphaleron}}} \providecommand{\unit}[1]{\ensuremath{\,\mathrm{#1}}} \providecommand{\eV}{\ensuremath{\,\mathrm{eV}}} \providecommand{\keV}{\ensuremath{\,\mathrm{keV}}} \providecommand{\MeV}{\ensuremath{\,\mathrm{MeV}}} \providecommand{\GeV}{\ensuremath{\,\mathrm{GeV}}} \providecommand{\TeV}{\ensuremath{\,\mathrm{TeV}}} \definecolor{blue}{rgb}{0,0,1} \definecolor{red}{rgb}{1,0,0} \definecolor{red}{rgb}{1,0,0} \definecolor{orange}{rgb}{1,0.4,0} \definecolor{green}{rgb}{0,1,0} \definecolor{grey}{rgb}{0.5,0.5,0.5} \providecommand{\eqn}{eqn.\,} \providecommand{\eqns}{eqns.\,} \providecommand{\Eqn}{Equation\,} \providecommand{\Eqns}{Equations\,} \providecommand{{\, .}}{{\, .}} \providecommand{{\, ,}}{{\, ,}} \providecommand{\, .}{\, .} \providecommand{\ifeqthenelse}[4]{\edef\tempa{#1}\def#2}\ifx\tempa\tempb {#3} \else {#4}\fi{#2}\ifx\tempa#2}\ifx\tempa\tempb {#3} \else {#4}\fi {#3} \else {#4}\fi} \providecommand{\formula}[2]{$#1\ifeqthenelse{#2}{}{}{\label{eqn:#2}}$} \providecommand{\bigformula}[2]{\begin{equation}#1\ifeqthenelse{#2}{}{\nonumber}{\label{eqn:#2}} \end{equation}} \providecommand{\bigformularray}[2]{\begin{eqnarray}#1\ifeqthenelse{#2}{}{\nonumber}{\label{eqn:#2}}\end{eqnarray}} \providecommand{\bigformulalign}[2]{\begin{align}#1\ifeqthenelse{#2}{}{\nonumber}{\label{eqn:#2}}\end{align}} \providecommand{\mref}[1]{eqn.\,(\ref{#1})} \providecommand{\eqnref}[1]{(\ref{eqn:#1})} \providecommand{\chap}{chapter~} \providecommand{\chapref}[1]{\ref{chap:#1}} \providecommand{\salign}{&&\hspace{-6mm}} \defGUT{GUT} \defQED{QED} \defQFT{QFT} \defC{C} \defCP{CP} \defCPT{CPT} \defB{B} \providecommand{\fig}{fig.\,} \providecommand{\Fig}{Fig.\,} \providecommand{\figref}[1]{\ref{fig:#1}} \providecommand{\app}{appendix\,} \providecommand{\appref}[1]{\ref{app:#1}} \providecommand{\sectref}[1]{\ref{sec:#1}} \providecommand{\sect}{section\,} \providecommand{\secref}[1]{\ref{sec:#1}} \providecommand{\dkn}[1]{#1} \providecommand{\todo}[1]{{\bf (TODO: #1)}} \definecolor{darkgreen}{rgb}{0.0,0.3,0.0} \newcommand{\edited}[1]{#1} \providecommand{\alt}[2]{{#1}\footnote{ALTERNATIVE: {#2}}} \providecommand{\note}[1]{{{\color{grey} #1}}} \providecommand{\order}[1]{\mathcal{O}(#1)} \providecommand{\floorbr}[1]{\lfloor #1\rfloor} \providecommand{\const}{\mathrm{const}} \providecommand{\Si}{\mathrm{Si}} \providecommand{\Ci}{\mathrm{Ci}} \providecommand{\sgn}{\mathrm{sgn}} \providecommand{\sig}{\sgn} \providecommand{\sign}{\sgn} \providecommand{\ramp}[1]{R#1} \providecommand{\abs}[1]{\left|#1\right|} \providecommand{\heaviside}[1]{\Theta\ifeqthenelse{#1}{}{}{\left(#1\right)}} \providecommand{\Min}[1]{\min\left(#1\right)} \providecommand{\Max}[1]{\max\left(#1\right)} \providecommand{\sbessel}[2]{j_{#1}(#2)} \providecommand{\partd}[2]{\frac{\partial #1}{\partial #2}} \providecommand{\partdiff}[3]{\left(\frac{\partial #1}{\partial #2}\right)_{#3}} \providecommand{\diff}[1]{{d}#1} \providecommand{\Dnm}{D^{nm}} \providecommand{\Dn}{D^{n}} \providecommand{\Dnind}[1]{D^{#1}} \providecommand{\Dnmind}[2]{D^{#1,#2}} \providecommand{\gaussbr}[1]{\left[#1\right]} \providecommand{\coeffa}[2]{a_{#1,#2}} \providecommand{\termc}[1]{\left( #1 \right)} \providecommand{\termtheta}{\Theta\left( k,p,q,r \right)} \providecommand{\termprefac}[1]{#1\termtheta} \providecommand{\funcU}[5]{U_{#1}\left(#2,#3,#4,#5\right)} \providecommand{\trash}[1]{} \providecommand{\intI}[8]{I(#1,#2,#3,#4;#5,#6,#7,#8)} \providecommand{\intIn}[9]{I(#1;#2,#3,#4,#5;#6,#7,#8,#9)} \providecommand{\intIthree}[6]{I(#1,#2,#3;#4,#5,#6)} \providecommand{\wignerthreej}[6]{ \begin{pmatrix} #1 & #2 & #3 \\ #4 & #5 & #6 \end{pmatrix} } \providecommand{\wignersixj}[6]{ \left\{ \begin{matrix} #1 & #2 & #3 \\ #4 & #5 & #6 \end{matrix} \right\} } \providecommand{\mathcal{M}}{\mathcal{M}} \providecommand{\mathscr{L}}{\mathscr{L}} \providecommand{\,d}{\,d} \providecommand{\chargec}[1]{({#1})^c} \providecommand{\bvec{k}}{{\bvec{k}}} \providecommand{\bvec{p}}{{\bvec{p}}} \providecommand{\bvec{q}}{{\bvec{q}}} \providecommand{\bvec{r}}{{\bvec{r}}} \providecommand{\bvec{\lambda}}{{\bm{\lambda}}} \providecommand{\bar{b}}{\bar{b}} \providecommand{\meavsqu}{\left|\mathcal{M}\right|^2} \providecommand{\meavsqudep}[1]{\left|\mathcal{M}\right|_{#1}^2} \providecommand{\amplitude}[2]{\mathcal{M}_{{#1},\,{#2}}} \providecommand{\ctkp}{\cos\left(\theta_{kp}\right)} \providecommand{\ctkq}{\cos\left(\theta_{kq}\right)} \providecommand{\ctkr}{\cos\left(\theta_{kr}\right)} \providecommand{\cttwovec}[2]{\cos\left(\theta_{\bvec{#1}\bvec{#2}}\right)} \providecommand{\ctonevec}[1]{\cos\theta_{\bvec{#1}}} \providecommand{\sttwovec}[2]{\sin\left(\theta_{\bvec{#1}\bvec{#2}}\right)} \providecommand{\stonevec}[1]{\sin\theta_{\bvec{#1}}} \providecommand{\ctq}{\cos\theta_{\bvec{q}}} \providecommand{\ctr}{\cos\theta_{\bvec{r}}} \providecommand{\ctl}{\cos\theta_{\bvec{\lambda}}} \providecommand{\stq}{\sin\theta_{\bvec{q}}} \providecommand{\str}{\sin\theta_{\bvec{r}}} \providecommand{\stl}{\sin\theta_{\bvec{\lambda}}} \newcommand{\f}[2]{f_{#2}^{#1}} \newcommand{\feq}[2]{f_{#2}^{#1,eq}} \newcommand{\flong}[3]{f^{#1}({#2},{#3})} \newcommand{\qstatf}[2]{(1-\xi^{#1}\f{#1}{#2})} \newcommand{\lhscollint}[4]{\frac{g_{#1}}{(2\pi)^3}\int\frac{\,d^3k}{\e{#1}{k}}\carray{{#2}\leftrightarrow {#3}}{k}{\f{#1}{},\f{#4}{}}} \newcommand{\abundance}[1]{Y_{#1}} \newcommand{\abundanceeq}[1]{Y^{eq}_{#1}} \defBoltzmann equation{Boltzmann equation} \defBoltzmann equations{Boltzmann equations} \defKadanoff-Baym equation{Kadanoff-Baym equation} \defKadanoff-Baym equations{Kadanoff-Baym equations} \defDirac-delta{Dirac-delta} \defDirac-deltas{Dirac-deltas} \newcommand{x}{x} \newcommand{\Delta k}{\Delta k} \newcommand{{G_F}}{{G_F}} \newcommand{{G_\rho}}{{G_\rho}} \providecommand{\indl}{n} \providecommand{\indk}{\alpha} \providecommand{\vecx}{\bvec{x}} \providecommand{\veck}{\bvec{k}} \providecommand{\veceta}{\bvec{\eta}} \providecommand{\vecetaunit}{{\mathbf{\hat\eta}}} \providecommand{\veckunit}{\hat{\bvec{k}}} \providecommand{\vecxunit}{\hat{\bvec{x}}} \providecommand{\vecxabs}{x} \providecommand{\veckabs}{k} \providecommand{\vecetaabs}{\eta} \providecommand{\laplace}{\Delta} \newcommand{k}{k} \newcommand{p}{p} \newcommand{q}{q} \newcommand{r}{r} \newcommand{\m}[1]{m_{#1}} \newcommand{\e}[2]{E^{#1}_{#2}} \newcommand{\pxi}[2]{\xi^{#1}_{#2}} \newcommand{\stat}[1]{\xi^{#1}} \newcommand{\n}[1]{n_{#1}} \newcommand{\nequ}[1]{n^{eq}_{#1}} \newcommand{\rho}{\rho} \newcommand{\liou}[2]{L_{#2}[#1]} \newcommand{\lioulong}[3]{\ifeqthenelse{}{#2}{\ifeqthenelse{}{#3}{L[#1]}{L[#1](#2,#3)}}{\ifeqthenelse{}{#3}{L[#1]}{L[#1](#2,#3)}}} \newcommand{\carray}[3]{C_{#2}^{#1}[#3]} \newcommand{\rate}[1]{\Gamma_{#1}} \def\eprinttmp@#1arXiv:#2 [#3]#4@{ \ifthenelse{\equal{#3}{x}}{\href{http://arxiv.org/abs/#1}{\texttt{#1}}}{\href{http://arxiv.org/abs/#1}{\texttt{#1}.}} } \newcommand{\eprint}[2]{\eprinttmp@#1arXiv: [y]@} \hyphenation{Deut-schen} \hyphenation{Astro-Teil-chen-phy-sik} \usepackage{color} \definecolor{blue}{rgb}{0,0,0.5} \definecolor{lightblue}{rgb}{0,0,1} \definecolor{red}{rgb}{1.0,0,0} \definecolor{lightred}{rgb}{1,0.5,0} \definecolor{green}{rgb}{0,0.5,0} \definecolor{darkgreen}{rgb}{0.0,0.3,0.0} \definecolor{grey}{rgb}{0.5,0.5,0.5} \renewcommand{\Im}{{\rm Im}} \providecommand{k}{{p}_1} \providecommand{p}{{p}_2} \providecommand{q}{{p}_3} \providecommand{r}{{p}_4} \providecommand{\momiabs}{|\bvec{p}_i|} \providecommand{\momabs}{|\bvec{p}|} \providecommand{\momkabs}{|\bvec{k}|} \providecommand{\mompabs}{|\bvec{p}|} \providecommand{\momqabs}{|\bvec{q}|} \providecommand{\momrabs}{|\bvec{r}|} \providecommand{\lorig}{L} \providecommand{\ltrafo}{{\tilde{L}}} \providecommand{\qstat}[1]{[1+#1]} \providecommand{\qstatferm}[1]{[1-#1]} \providecommand{\feqtrans}[1]{f^{b,eq}_a\ifeqthenelse{}{#1}{}{(#1)}} \providecommand{\fbareqtrans}[1]{f^{\bar{b} ,eq}_a\ifeqthenelse{}{#1}{}{(#1)}} \providecommand{\hubblerate}{{H}} \newcommand{\carraykb}[3]{\tilde{C}_{\ifeqthenelse{}{#2}{}{#2}}^{#1}\ifeqthenelse{}{#3}{}{[{#3}]}} \newcommand{\carrayorigkb}[3]{{C}_{#1}\ifeqthenelse{}{#3}{}{[{#3}]}{\ifeqthenelse{}{#2}{}{(#2)}}} \newcommand{M_i}{M_i} \newcommand{M_j}{M_j} \begin{document} \pacs{11.10.Wx, 98.80.Cq} \keywords{Kadanoff--Baym equations, Boltzmann equation, expanding universe, leptogenesis, thermal quantum field theory} \preprint{TUM-HEP-749/10} \title{Medium corrections to the \textit{CP}-violating parameter in leptogenesis } \author{M. Garny$^{b}$} \email[\,]{<EMAIL>} \author{A. Hohenegger$^{a}$} \email[\,]{<EMAIL>} \author{A. Kartavtsev$^{a}$} \email[\,]{<EMAIL>} \affiliation{% \vskip 2mm $^a$Max-Planck-Institut f\"ur Kernphysik, Saupfercheckweg 1, 69117 Heidelberg, Germany\\ $^b$Technische Universit\"at M\"unchen, James-Franck-Stra\ss e, 85748 Garching, Germany} \begin{abstract} In two recent papers, arXiv:0909.1559 and arXiv:0911.4122, it has been demonstrated that one can obtain quantum corrected Boltzmann kinetic equations for leptogenesis using a top-down approach based on the Schwinger--Keldysh/Kadanoff--Baym formalism. These ``Boltzmann-like'' equations are similar to the ones obtained in the conventional bottom-up approach but differ in important details. In particular there is a discrepancy between the CP-violating parameter obtained in the first-principle derivation and in the framework of thermal field theory. Here we demonstrate that the two approaches can be reconciled if \textit{causal} $n$-point functions are used in the thermal field theory approach. The new result for the medium correction to the CP-violating parameter is \textit{qualitatively} different from the conventional one. The analogy to a toy model considered earlier enables us to write down consistent quantum corrected Boltzmann equations for thermal leptogenesis in the SM+$3\nu_R$ which include quantum statistical terms and medium corrected expressions for the CP-violating parameter. \end{abstract} \maketitle \newcommand{\cpviol}[3]{\epsilon_{#1}^{#2,#3}} \section{Introduction} To calculate the baryon asymmetry generated during the epoch of leptogenesis \cite{Fukugita:1986hr} in the standard model extended by three right-handed neutrinos (SM+$3\nu_R$) and its extensions one usually uses standard Boltzmann kinetic equations. The collision terms (and in particular the CP-violating parameters) in this equations are computed in vacuum in the in-out formalism \cite{Giudice:2004npb685,Davidson:pr2008} and do not take into account effects induced by the hot medium of the early universe. Such effects can be consistently taken into account in a top-down approach based on the Schwinger--Keldysh/Kadanoff--Baym formalism. In \cite{Garny:2009rv,Garny:2009qn} we have applied it to a simple toy model of leptogenesis and derived a new (quantum corrected) form of the Boltzmann equations, which includes quantum statistical factors and takes the medium effects into account. We have found that the medium corrections to the CP-violating parameter $\epsilon$ depend only linearly on the one-particle distribution functions (see also \cite{Kiessig:2009cm,Buchmuller:2000nd,Anisimov:2010aq}). In the analysis based on finite temperature field theory for the phenomenological scenario of thermal leptogenesis \cite{Giudice:2004npb685,Davidson:pr2008,Covi:1998prd57} and for GUT baryogenesis \cite{Takahashi:1984} the medium corrections to the CP-violating parameter $\epsilon$ depend quadratically on the distribution functions. This discrepancy has been noted in the context of leptogenesis in \cite{Garny:2009rv} for the vertex contribution to the CP-violating parameter and later in \cite{Garny:2009qn} for the self-energy contribution. Here, we use a finite temperature equivalent of the Cutkosky cutting rules \cite{Kobes1985np260p3,Kobes:1986npb272,Bedaque:1996mpl,Gelis:1997npb508} to derive thermal corrections to the expression for the imaginary part of the three-point vertex function and the self-energy loop and to calculate the corresponding medium-corrected CP-violating parameters. We show that the discrepancy is due to an ambiguity in the real-time (RTF) formulation of thermal quantum field theory and disappears if one considers \textit{retarded} or \textit{advanced} $n$-point functions. In the framework of the toy model this has been demonstrated recently in \cite{Hohenegger:2009phd}. Together with the new form of the Boltzmann equation derived in \cite{Garny:2009rv,Garny:2009qn} this puts us in the position to write down quantum corrected {Boltzmann equations} for the phenomenological scenario of thermal leptogenesis which consistently include the medium corrected CP-violating parameter and quantum statistical terms. In section \ref{sec:TQFT}, we introduce our notations for the CP-violating parameters and the thermal field theory formalism. Then, in section \ref{PhysGhost} we review the conventional calculation of the thermal corrections, and in section \ref{Causal} we demonstrate how to reconcile them with the recent results from nonequilibrium field theory. Finally, in section \ref{BoltzmannEq} we present the quantum corrected Boltzmann equations taking medium corrections into account. \section{\label{sec:TQFT}CP-violating parameter and thermal field theory} In the phenomenological scenario of thermal leptogenesis as well as in the toy model, the matter-antimatter asymmetry is generated by the decay of a heavy species. In both cases the {CP} violation in this decay is caused by interference between the tree level and the one-loop diagrams, see \fig\ref{fig:majo_decays}. \begin{figure}[th] \centering \includegraphics{figures/fmf_thermal_leptogenesis_one_loop_corrections} \caption{Tree level and one-loop contributions to the heavy Majorana neutrino decay $\psi_i\rightarrow \alpha\beta$. The asymmetry, at lowest order, is due to the interference of these contributions.\label{fig:majo_decays}} \end{figure} In the phenomenological scenario $\psi_i ={N}_i$ are heavy Majorana neutrinos which decay via Yukawa interactions ${\cal L}=h_{\alpha i}\ell_\alpha N_i\, \phi+h.c.$ into leptons $\alpha ={\ell}$ and Higgs $\beta =\phi$ or their anti-particles. In the toy model $\psi_i$ is a heavy real scalar particle which decays via Yukawa interactions ${\cal L}=-\frac{g_i}{2!}\psi_i bb+h.c.$ into two light scalars $\alpha =\beta =b$ or the conjugate $\bar{b}$. The CP-violating parameter $\epsilon_i$ for the decay of $\psi_i$ is defined as \bigformulalign{ \epsilon_i&=\epsilon_i^V +\epsilon_i^S =\frac{\Gamma_{\psi_i\rightarrow \alpha\beta} -\Gamma_{\psi_i\rightarrow\bar{\alpha}\bar{\beta}}}{\Gamma_{\psi_i\rightarrow \alpha\beta}+\Gamma_{\psi_i\rightarrow\bar{\alpha}\bar{\beta}}}{\, ,} }{CP-violating parameter} where $\Gamma_{N_i\rightarrow {{\ell}}{\phi}}$ includes a sum over flavour indices and loop-internal Majorana neutrino generations in the case of the phenomenological scenario: $\Gamma_{N_i\rightarrow {{\ell}}{\phi}}=\sum_{\alpha,\,j} \Gamma_{N_i\rightarrow {{\ell}_\alpha}{\phi}}$ (we do not consider flavor effects here). If the tree level and one-loop contributions are written as $\lambda_0 \mathcal{A}_0$ and $\lambda_1 \mathcal{A}_1$, respectively, where all coupling constants are absorbed in $\lambda_{0(1)}$, the CP-violating parameter becomes at lowest order: \bigformula{ \epsilon_i=\frac{\abs{\lambda_0 \mathcal{A}_0+\lambda_1\mathcal{A}_1}^2-\abs{\lambda_0^*\mathcal{A} _0+\lambda_1^*\mathcal{A}_1}^2}{\abs{\lambda_0\mathcal{A}_0+\lambda_1\mathcal{A} _1}^2+\abs{\lambda_0^*\mathcal{A}_0+\lambda_1^*\mathcal{A}_1}^2}\simeq -2\frac{\Im\big\{\lambda_0^*\lambda_1\big\}\Im\left\{\mathcal{A}_0^*\mathcal{A} _1\right\}}{{\abs{\lambda_0}^2}\abs{\mathcal{A}_0}^2}{\, ,} }{cp asymmetry in terms of imaginary parts}where the sum over lepton flavour and Majorana neutrino generation indices is again implicit. In the case of thermal leptogenesis this leads to \bigformula{ \epsilon_i=-2\sum_{j\neq i}\frac{\Im\big\{(h^{\dagger}h)_{ij}^2\big\}}{(h^{\dagger}h)_{ii}}\frac{ \Im\left\{\mathcal{A}_0^*\mathcal{A}_1\right\}}{2\, q\cdot k}{\, ,}\quad i=1,2,3{\, ,} }{cp asymmetry in terms of imaginary parts phenomenological} where $q$ and $k$ denote the four-momenta of the Majorana neutrino and the lepton respectively, see \fig\figref{vertex corrections momentum flow}, and for the toy model to \bigformula{ \epsilon_i=-2\abs{g_j}^2 \Im\bigg(\frac{g_i g_j^*}{g_i^* g_j}\bigg)\Im\left\{\mathcal{A}_0^* \mathcal{A}_1\right\}{\, ,}\quad i\neq j {\, ,} \,\quad i=1,2{\, .} }{cp asymmetry in terms of imaginary parts toy} This means that one needs to compute the imaginary (absorptive) part of the vertex and the self-energy loop contributions $\Im\left\{\mathcal{A}_0^*\mathcal{A}_1\right\}$. In vacuum this can be done conveniently with help of the Cutkosky cutting rules \cite{cutkosky:429,Eden:2002,Bellac:1992}. In thermal quantum field theory these can be generalized in order to take into account interactions of internal lines in the loops with the background medium \cite{Kobes1985np260p3,Bedaque:1996mpl,Gelis:1997npb508}. In the real-time formalism of thermal quantum field theory two types of fields, termed {\em type-1} and {\em type-2} fields, are introduced in order to avoid pathological singularities \cite{Rivers:1988}. Vertices can be of either type, differing only by a relative minus sign. We denote them by $g^1 = -ig$ and $g^2 = +ig$ for a generic coupling\footnote{At the end of the calculation of $\Im\left\{\mathcal{A}_0^*\mathcal{A}_1\right\}$ we set $g=1$, since the physical coupling constants have been factored out into $\lambda_0$ and $\lambda_1$.} $g$. The propagators connecting the different types of vertices can be considered as components of a $2\times 2$ propagator matrix\footnote{In \cite{Giudice:2004npb685} and elsewhere resummed propagators have been used in this place to prevent the appearance of singularities. Since we are mainly interested in the structure of the thermal corrections we stick to the free thermal propagators here.} \bigformula{ G_{a}(p)= \left(\begin{array}{cc} G_a^{11}(p) & G_a^{12}(p)\\ G_a^{21}(p) & G_a^{22}(p) \end{array}\right) = \left(\begin{array}{cc} \Delta_a (p) & e^{\beta p_0 /2}\Delta_a^- (p) \\ e^{-\beta p_0 /2}\Delta_a^+ (p) & \Delta_a^* (p) \end{array}\right) {\, .}}{} For a scalar particle $b$ the components are \bigformulalign{ \Delta_b (p) = D_b (p){\, ,}\quad \Delta_b^{\pm} (p) = D_b^{\pm}(p){\, .} }{thermal Feynman rules scalars} For a fermion $f$ the components are \bigformulalign{ \Delta_f (p) = (\gamma\cdot p +\m{f}) D_f (p){\, ,}\quad \Delta_f^{\pm} (p) =(\gamma\cdot p +\m{f}) D_f^{\pm}(p){\, .} }{thermal Feynman rules fermions} For brevity we have defined \bigformulalign{ D_a (p) & = \frac{i}{p^2 -\m{a}^2 +i\epsilon} - 2\pi \xi_a \feq{a}{}(p)\delta{(p^2 -\m{a}^2)}{\, ,}\nonumber\\ D_a^{\pm}(p) & = 2\pi \left[\heaviside{\pm p_0} - \xi_a \feq{a}{}(p)\right]\delta (p^2 -\m{a}^2){\, ,} }{thermal Feynman rules D} where $\xi_a = +1$ for fermions and $\xi_a = -1$ for bosons. Here, we denote by $\feq{b}{}(p)$ and $\feq{f}{}(p)$ the equilibrium distribution function for bosons and fermions, respectively, given by \begin{equation} \feq{a}{}(p) = \left[ \exp\left(\beta |p_\mu U^\mu| \right) + \xi_a \right]^{-1} \;. \end{equation} They are functions of the Lorentz invariant product $p_\mu U^\mu$ of the particles' four-momentum and the four-velocity $U$ of the plasma in a general frame. In the rest-frame of the plasma, $U=(1,0,0,0)$, we obtain the standard form which depends on $p_0$. In the following we assume that it is sufficient to replace the different propagators in our toy model and the phenomenological theory by their thermal field theory equivalents given in \eqn\eqnref{thermal Feynman rules scalars}. This approach has been followed in previous works for the baryogenesis and leptogenesis scenarios \cite{Giudice:2004npb685,Davidson:pr2008,Covi:1998prd57,Takahashi:1984}. We ignore further thermal effects, such as thermal corrections to the masses and wave function renormalization here. Denoting vertices attached to external lines by $x_i$ and those attached to internal lines only by $z_j$ we can formally denote an amputated $n$-point graph by $F(x_1,\ldots ,x_n;z_j)$. Here we assume that $F$ is given in momentum space, writing the position space coordinates in order to identify the individual vertices. The contribution of this graph to the amplitude is $-iF(x_1,\ldots ,x_n;z_j)$. Physical amplitudes involve a sum over possible combinations of types of internal vertices: \bigformula{ \mathcal{F}(x_1,\ldots ,x_n;z_j) = \sum_{\text{type}\,z_j} F(x_1,\ldots ,x_n;z_j){\, .} }{} For external vertices of fixed type it has been shown \cite{Kobes1985np260p3,Kobes:1986npb272} that this sum is equivalent to a sum over all possible ``circlings'' of the internal vertices:\footnote{The historic origin of this formula was that the external fields where considered to be all of type 1 (physical).} \bigformulalign{ \mathcal{F}(x_1,\ldots ,x_n;z_j) = & \sum_{\text{circling}\, z_j} F_{\gtrless}(x_1,\ldots ,x_n;z_j){\, .} }{circling with fixed type of external vertices} $F_{>}$ and $F_{<}$ with ``circled'' vertices represent graphs computed using the set of rules, given in \fig\figref{circling rules}. These differ for the computation of $F_{>}$ and $F_{<}$ by interchange of the $\Delta^+$ and $\Delta^-$ propagators. In $F_{\gtrless}(x_1,\ldots ,x_n;z_j)$ we explicitly denote circling of a vertex $\alpha$ as $F_{\gtrless}(x_1,\ldots ,\underline{x}_{\alpha},\ldots ,x_n;z_j)$. Note that the two ways of defining $\mathcal{F}$ in terms of $F_{>}$ and $F_{<}$ in \eqn\eqnref{circling with fixed type of external vertices} are in agreement only if the Kubo--Martin--Schwinger (KMS) boundary condition, \begin{equation} \Delta_a^- (p) = -\xi_a e^{ - \beta p\cdot U}\Delta_a^+ (p) \;, \end{equation} is satisfied. This is the case in thermal equilibrium. \begin{figure}[!ht] \centering \includegraphics{figures/fmf_vertex_circled}\\ \vspace{10mm} \includegraphics{figures/fmf_propagator_circled} \caption{Circling rules for a generic theory used for the computation of $F_{>}$ in momentum space. The rules for the computation of $F_{<}$ differ by interchange of $\Delta^+ (p)$ and $\Delta^- (p)$. The $\Delta^{\pm}$ propagators connecting circled and uncircled vertices may be interpreted as cut propagators. In vacuum they correspond to the cut propagators in the Cutkosky rules.\label{fig:circling rules}} \end{figure} From $\mathcal{F}$ we can then compute $\Im\{\mathcal{A}_0^* \mathcal{A}_1\}$ as\footnote{In the phenomenological scenario the Feynman rules for Majorana neutrinos include spinors, charge conjugation and projection operators which we assume to be included in $F_{\gtrless}$.} \bigformula{ \Im\{\mathcal{A}_0^* \mathcal{A}_1\} = - \Im\bigg\{\frac{i^{-1}\mathcal{F}}{g_1 g_2 g_3}\bigg\}{\, .} }{contribution to absorptive part of amplitude} where $g_1$, $g_2$ and $g_3$ stand for the generic couplings associated with the three vertices in the one-loop diagrams at Fig. \ref{fig:circling rules}. \section{\label{PhysGhost}Physical and ghost fields} In this section we briefly review the conventional calculation of the CP-violating parameter in real-time thermal field theory. However, we use a notation that is helpful to understand the ambiguities emerging there, and that can be more easily compared to the results from non-equilibrium field theory. An obvious problem with the real-time formulation for the computation of $n$-point functions is that there are in general $2^n$ such functions which differ in the types of the external vertices. Historically the correct function was considered to be the one with all external vertices of type-1 (physical). In this case \eqn\eqnref{circling with fixed type of external vertices} leads to the following formula for the imaginary part of a graph's contribution to the amplitude: \bigformula{ \Im\left\{i^{-1}\mathcal{F}(1,\ldots ,1;z_j)\right\} =\frac{1}{2}\sum_{\text{circling}\,(x_i),\, z_j}F_{\gtrless}(x_1 ,\ldots ,x_n;z_j){\, ,} }{circling formula by Kobes Semenoff} where the sum includes all possible circlings of the internal vertices $z_j$ but only those circlings of external vertices $x_i$ which include both, circled and uncircled vertices (indicated by the brackets around $x_i$). \begin{figure}[!ht] \centering \includegraphics{figures/fmf_vertex_corrections_momentum_alow} \caption{Momentum flow in the vertex and the self-energy loop.\label{fig:vertex corrections momentum flow}} \end{figure} The six diagrams contributing to the imaginary part of the three-point vertex function according to \eqn\eqnref{circling formula by Kobes Semenoff} are shown in \fig\figref{vertex corrections imaginary part by Kobes Semenoff}. \begin{figure}[!ht] \centering \includegraphics{figures/fmf_vertex_corrections_ainite_temperature} \caption{Circlings contributing to $\Im\left\{i^{-1}\mathcal{F}(1,1,1)\right\}$ for the vertex loop. At one-loop level the circlings can be interpreted as cuts, as indicated, by the lines separating circled from uncircled regions \cite{Kobes1985np260p3}. The contributions from diagrams involving cuts through the $x_2$-$x_3$ line are suppressed relative to the others in the hierarchical limit.\label{fig:vertex corrections imaginary part by Kobes Semenoff}} \end{figure} The circlings contributing to the self-energy part are shown in \fig\figref{selfenergy loop corrections imaginary part by Kobes Semenoff}. \begin{figure}[!h] \centering \includegraphics{figures/fmf_loop_corrections_ainite_temperature} \caption{Circlings contributing to $\Im\left\{i^{-1}\mathcal{F}(1,1;z)\right\}$ for the self-energy loop. The graphs (b) and (c) vanish since $\psi_i$ and $\psi_j$ cannot be on-shell simultaneously. Note that we consider only the diagrams with $\psi_i$ in the external and $\psi_j$ in the internal line ($i\neq j$) because these are the only ones which contribute to $\epsilon_i$.\label{fig:selfenergy loop corrections imaginary part by Kobes Semenoff}} \end{figure} The contributions which correspond to cuts through the $\psi_j$ line are suppressed in the hierarchical limit. If they are neglected the application of this circling formula leads for the toy model to the result \bigformulalign{ \cpviol{i}{V}{th}=-\frac1{8\pi}\frac{|g_j|^2}{M_i^2}\Im\left(\frac{g_i g_j^*}{g_i^* g_j}\right)\int\frac{\,d\Omega_l}{4\pi}\frac{1+\feq{\bar{b}}{E_1}+\feq{\bar{b}}{E_2 }+2\feq{\bar{b}}{E_1}\feq{\bar{b}}{E_2}}{ M_j^2/M_i^2+\frac{1}{2}(1+\cos\theta_l)}+\ldots{\, ,} }{epsilon vertex thermal} for the vertex contribution and \bigformulalign{ \cpviol{i}{S}{th}=-\frac{\abs{g_j}^2}{16\pi}\Im\left(\frac{g_i g_j^*}{g_i^* g_j}\right)\frac{1}{M_j^2 -M_i^2}\int\frac{\,d\Omega_l}{4\pi}\big\{{1+\feq{\bar{b}}{E_1}+\feq{\bar{b}}{E_2 }+2\feq{\bar{b}}{E_1}\feq{\bar{b}}{E_2}}\big\} }{epsilon selfenergy loop thermal} for the self-energy contribution. The distribution functions are to be evaluated for the energies $E_{1}$ and $E_{2}$ given by\footnote{Note that if in \eqn\eqnref{epsilon selfenergy loop thermal} the term quadratic in the distribution functions was absent, then by redefining the integration variable $\varphi_l$ we could write the energies $E_1$ and $E_2$ in the form $E_{{1,2}}={\frac12}\big[\e{\psi_1}{q} +\momqabs(\sin\theta_l\cos\varphi_l\cos\delta' \mp \cos\theta_l\sin\delta')\big]$, which was used in \cite{Hohenegger:2009a}.} \begin{align} \label{IntermedEnergy} E_{{1,2}}={\frac12}\big[\e{\psi_1}{q} \mp\momqabs(\sin\theta_l\cos\varphi_l\cos\delta' + \cos\theta_l\sin\delta')\big]{\, ,} \end{align} where $\theta_l$ and $\varphi_l$ are elements of the solid angle $\Omega_l$ and the angle $\delta'$ is given in the limit of massless decay products by $\sin\delta'=(\mompabs -\momkabs)/\momqabs$. The dots in \eqn\eqnref{epsilon vertex thermal} represent further terms in $\feq{\psi_j}{}$ which are neglected. Equivalently these results can be derived directly using only the $11$ components of the propagators, because the vertex and the self-energy loop do not include internal vertices. Very similar results are known for the phenomenological scenario which can be obtained using the propagators in \eqn\eqnref{thermal Feynman rules fermions} for fermions for the Majorana neutrinos and leptons in the loops and \eqn\eqnref{thermal Feynman rules scalars} for the Higgs bosons \cite{Giudice:2004npb685,Covi:1998prd57}. For the dependence on the distribution functions one obtains then the quadratic form \bigformula{ 1-\feq{{\ell}}{E_1}+\feq{\phi}{E_2}-2\feq{{\ell}}{E_1}\feq{\phi}{E_2} {\, .} }{thermal corrections to epsilon phenomenological scenario old} The results obtained from non-equilibrium field theory in \cite{Garny:2009rv,Garny:2009qn} differ from \eqns\eqnref{epsilon vertex thermal} and \eqnref{epsilon selfenergy loop thermal}. The non-equilibrium results feature a different dependence on the distribution functions, \bigformula{ {1+\f{\bar{b}}{E_1}+\f{\bar{b}}{E_2}+2\f{\bar{b}}{E_1}\f{\bar{b}}{E_2} \quad \rightarrow \quad 1+\f{\bar{b}}{E_1}+\f{\bar{b}}{E_2}}{\, .} }{distribution dependence toy model} Note that the top-down results are valid even if $\f{\bar{b}}{}$ is not an equilibrium distribution ($\f{\bar{b}}{}\simeq \f{b}{}$ must hold, however) and that the dependence is linear in the distribution function in contrast to \eqns\eqnref{epsilon vertex thermal}, \eqnref{epsilon selfenergy loop thermal}. The latter property contradicts the result derived from thermal quantum field theory . In the phenomenological model, an analogous replacement leads to a particularly important discrepancy. Indeed, \eqn\eqnref{thermal corrections to epsilon phenomenological scenario old} would imply a cancellation of the leading effects since $\feq{\phi}{p}-\feq{{\ell}}{p}=2\feq{\phi}{p}\feq{{\ell}}{p}$. The remaining effect is, in this case, entirely due to the fact that different energies enter the distribution functions of leptons and Higgs particles in \eqn\eqnref{thermal corrections to epsilon phenomenological scenario old}. Since $E_1-E_2\sim \momqabs$, this effect vanishes when the velocity of the Majorana neutrino in the medium rest-frame, $\momqabs/\e{N_1}{q}$, becomes small. Therefore it is important to check wether a replacement of the form of \eqn\eqnref{distribution dependence toy model} does also occur in the phenomenological scenario. This will be investigated in the next section. \section{\label{Causal}Causal n-point functions} We will now see how the finite temperature field theory approach can be reconciled with the results derived from non-equilibrium quantum field theory. In \cite{Kobes1990prd42,Kobes1991prd43,Eijck199plb278} it was shown that the combination \bigformulalign{ \mathcal{F}_{R/A}^{(\alpha)}(x_1,\ldots ,x_n;z_j) = & \sum_{\text{circling}\, x_i ,z_j}^{i\neq \alpha} F_{\gtrless}(x_1,\ldots x_\alpha ,\ldots ,x_n;z_j){\, ,} }{retarded product} referred to as the retarded (advanced) product, has the distinguishing property that the time component $(x_\alpha)_0$ is singled out as being the largest (smallest). This becomes clear when we consider the so-called largest (smallest) time equation \bigformula{ F_{\gtrless} (x_1,\ldots ,x_\alpha ,\ldots ,x_n) + F_{\gtrless} (x_1,\ldots ,\underline{x}_\alpha ,\ldots ,x_n) = 0\,,\quad\text{if}\,\,(x_\alpha)_0\,\,\text{largest/smallest}{\, ,} }{largest smallest time equation} which implies pairwise cancellation of the terms in \eqn\eqnref{retarded product} if any external vertex $x_i$ with $i\neq\alpha$ has the largest (smallest) time component. It has been realized that such causal products appear in Boltzmann equations in different cases, see for example \cite{Kobes1991prd43, Weldon:1983}. Furthermore, it has been shown that the causal products agree with the results of the calculation in imaginary-time formalism analytically continued to real energies, at least in a few examples including the self-energy loop and the three-point vertex. The imaginary part of the causal sum was shown in \cite{Kobes1991prd43} to obey \bigformulalign{ \Im\big\{i^{-1}\mathcal{F}_{R/A}^{(\alpha)}(x_1,\ldots ,x_\alpha ,\ldots , x_n;z_j)\big\} &=\nonumber\\ \mp\frac{1}{2}\sum_{\text{circling}\, x_i}^{\text{not all}}\sum_{\text{circling}\,z_j}^{}\Im\Big\{&i^{-1}F_> (x_1,\ldots ,\underline{x}_\alpha ,\ldots , x_n;z_j) - \nonumber\\ - & i^{-1}F_< (x_1,\ldots ,\underline{x}_\alpha ,\ldots , x_n;z_j)\Big\}{\, ,} }{imaginary part of causal products} where ``not all'' means that not all $x_i$ should be circled at the same time and the imaginary part is taken of the causal product in momentum space. Here, the vertex $x_\alpha$ with largest or smallest time is always circled. We can now compute the imaginary part of the advanced product $\Im\big\{i^{-1}\mathcal{F}_{A}^{(1)}(x_1,x_2,x_3)\big\}$ for the three-point vertex with smallest time component $(x_1)_0$ of the decaying particle. The relevant circlings are shown in \fig\figref{vertex corrections imaginary part causal products}. As before the contributions \fig\figref{vertex corrections imaginary part causal products}(b) and (c) are suppressed due to the cut through the $\psi_j$ propagator line. \begin{figure}[!ht] \centering \includegraphics{figures/fmf_vertex_corrections_ainite_temperature_causal} \caption{Circlings contributing to $\Im\big\{i^{-1}\mathcal{F}_{A}^{(1)}(x_1, x_2 ,x_3 )\big\}$ for the vertex loop. The advanced three-point function involves a difference of $F_>$ and $F_<$ contributions which differ by the replacement $\Delta^{\pm}\leftrightarrow\Delta^{\mp}$ in the circling rules. Since the finite temperature contributions (terms proportional to $f^{eq}$) to the latter are the same, all contributions quadratic in the distribution functions cancel. \label{fig:vertex corrections imaginary part causal products}} \end{figure} \newcommand{P_L}{P_L} \newcommand{P_R}{P_R} We compute the remaining contribution from \fig\figref{vertex corrections imaginary part causal products}(a): \bigformulalign{ \Im\big\{i^{-1}\mathcal{F}_{A}^{(1)}(x_1, x_2 ,x_3)\big\}=&\frac{1}{2}\Im\big\{i^{-1}F_{>}(\underline{x}_1,{x}_2,{x}_3)-i^{-1} F_{<}(\underline{x}_1,{x}_2,{x}_3)\big\}\nonumber\\ = & \frac{1}{2} \Im\Big\{i^{-1} \int \frac{\,d^4 l}{(2\pi)^4} \Big[(+ig_1)(-ig_2)(-ig_3)D_{\beta}^- (q+l)D_{\psi_j} (q+l-k)D_{\alpha}^+ (l) -\nonumber\\ & - (+ig_1)(-ig_2)(-ig_3)D_{\beta}^+ (q+l)D_{\psi_j} (q+l-k)D_{\alpha}^- (l) \Big] S\Big\}{\, ,} }{causal circling formula contribution a 1} where we take $F_{\gtrless}$ to include the spinors and charge conjugation operators $C$ as well as projection operators $P_R,\,P_L$ associated with the vertices (for the Majorana neutrino interactions). This leads to the trace part denoted by $S$ \cite{Buchmuller:1997yu,Giudice:2004npb685,Covi:1996wh}. In the massless lepton and Higgs limit: \bigformulalign{ S=&\sum_{\text{spins}}\big[\bar{u}_{{\ell}}(k)P_L u_{N_i}(q)\big]^*\big[\bar{u}_{{\ell}}(k)P_L (\gamma\cdot (q+l-k) +M_j)C^{-1} P_L (\gamma\cdot l) P_RC u_{N_i}(q)\big]\nonumber\\ = &\tr\big[(\gamma\cdot q -M_i)(\gamma\cdot k)M_j (\gamma\cdot l)P_R\big] \nonumber\\ =&-2\,M_i M_j\, k\cdot l }{trace part phenomenological} and $S=1$ for the toy model. It turns out that the pole of the $\psi_j$ propagator does not lie in the loop integration region, so we can drop the $i\epsilon$ prescription. We then get (the upper and lower signs correspond to the toy model and phenomenological scenario respectively) \bigformulalign{ \Im\big\{i^{-1}\mathcal{F}_{A}^{(1)}(x_1, x_2 ,x_3)\big\}=-\frac{1}{2}\Im\int \frac{\,d^4 l}{(2\pi)^2} & \delta\big((q+l)^2 -\m{\beta}^2\big)\delta\big(l^2-\m{\alpha}^2\big)\frac{i}{(q+l-k)^2 -M_j^2}\times\nonumber\\ \Big\{&\Big[\heaviside{-(q_0 +l_0)}\heaviside{l_0}-\heaviside{q_0 +l_0}\heaviside{-l_0}\Big]\nonumber\\ \pm&\Big[\heaviside{-(q_0 +l_0)}-\heaviside{q_0 +l_0}\Big]\feq{\alpha}{l}+\nonumber\\ +&\Big[\heaviside{l_0}-\heaviside{-l_0}\Big]\feq{\beta}{q+l}\Big\}S{\, ,} }{causal circling formula contribution a 4} which becomes \bigformulalign{ \Im\big\{i^{-1}\mathcal{F}_{A}^{(1)}(x_1, x_2 ,x_3)\big\}=\nonumber\\ =\frac{1}{2}\int \frac{\,d^4 l}{(2\pi)^2} \delta\big((q+l)^2 & -\m{\beta}^2\big)\delta\big(l^2-\m{\alpha}^2\big)\frac{1}{(q+l-k)^2 -M_j^2}\Big\{1\pm\feq{\alpha}{l}+\feq{\beta}{q+l}\Big\}S{\, .} }{causal circling formula contribution a 5} Performing the integration over $\,d \abs{\bvec{l}}\,d l_0$, this indeed leads to the result for the CP-violating parameter obtained in the top-down approach with correct dependence on the distribution functions \eqn\eqnref{distribution dependence toy model}. For the phenomenological scenario we obtain in the limit of massless lepton and Higgs: \bigformulalign{ \epsilon_i^{V,th} &=\frac{1}{16\pi}\sum_{j\neq i}\frac{{\Im}\left\{(h^\dagger h)_{ij}^2\right\}}{(h^\dagger h)_{ii}}\frac{M_j}{M_i}\int\frac{\,d\Omega_l}{4\pi}\frac{1-\cos\theta_l}{{M_j^2 }/{M_i^2}+\frac{1}{2}(1+\cos\theta_l)}\big\{1-\feq{{\ell}}{E_1}+\feq{\phi}{ E_2}\big\}{\, ,} }{CP-violating parameter thermal vertex} where $E_{1,2}$ are given by \eqn\eqref{IntermedEnergy}. In the zero temperature limit this reduces to the well-known result \bigformulalign{ \epsilon_i^{V,vac}&=-\frac{1}{8\pi}\sum_{j\neq i}\frac{{\Im}\left\{(h^\dagger h)_{ij}^2\right\}}{(h^\dagger h)_{ii}}f\left(\frac{M_j^2}{M_i^2}\right){\, ,} }{} with \begin{align} f(x)&=\sqrt{x}\left[1- (1+x)\ln\left(\frac{1+x}{x}\right)\right]{\, .} \end{align} The same computation can be performed for the self-energy loop. The possible circlings are shown in \fig\figref{selfenergy loop corrections imaginary part causal products}: \begin{figure}[!ht] \centering \includegraphics{figures/fmf_loop_corrections_ainite_temperature_causal} \caption{Circlings contributing to $\Im\big\{i^{-1}\mathcal{F}_{A}^{(1)}(x_1, x_2 ; z)\big\}$ for the self-energy loop. Graph (b) vanishes since $\psi_i$ and $\psi_j$ cannot be on-shell simultaneously.\label{fig:selfenergy loop corrections imaginary part causal products}} \end{figure} \bigformulalign{ \Im\big\{i^{-1}\mathcal{F}_{A}^{(1)}(x_1, x_2 ; z)\big\}=&\frac{1}{2}\Im\big\{i^{-1} F_{>}(\underline{x}_1,{x}_2,{x}_3)- i^{-1}F_{<}(\underline{x}_1,{x}_2,z)\big\}\nonumber\\ =& -\frac{1}{2} \Im\Big\{ \int\frac{\,d^4 l}{(2\pi)^4} \Big[(+ig_1)(-ig_z)(-ig_3)D_{\beta}^- (q+l)D_{\psi_j} (q)D_{\alpha}^+ (l) -\nonumber\\ &\hspace{20mm} - (+ig_1)(-ig_z)(-ig_3)D_{\beta}^+ (q+l)D_{\psi_j} (q)D_{\alpha}^- (l) \Big] S\Big\}{\, ,} }{causal circling formula contribution b 1} where (the result for) $S$ coincides with \eqn\eqnref{trace part phenomenological} in the phenomenological scenario, while $S=1/2!$ includes an additional symmetrization factor in the toy model. This becomes \bigformulalign{ \Im\big\{i^{-1}\mathcal{F}_{A}^{(1)}(x_1, x_2 ; z)\big\}=\nonumber\\ =\frac{1}{2}\int \frac{\,d^4 l}{(2\pi)^2} & \delta\big((q+l)^2 -\m{\beta}^2\big)\delta\big(l^2-\m{\alpha}^2\big)\frac{1}{q^2 -M_j^2}\Big\{1\pm\feq{\alpha}{l}+\feq{\beta}{q+l}\Big\}S{\, .} }{causal circling formula contribution b 5} This corresponds to the result for the self-energy contribution in the hierarchical limit in the top-down approach \cite{Garny:2009qn} if the equilibrium distribution functions are replaced with non-equilibrium ones. In the zero temperature limit this leads to the correct vacuum result. Thus, we have shown that (within the toy model) the CP-violating parameter $\epsilon^{th}$ obtained with help of thermal quantum field theory coincides with the one obtained in the top-down approach (in the approximately symmetric case) when one uses causal products instead of the conventional ones which assume type-1 external vertices. Furthermore, by comparing with the top-down result, we find that the thermal field theory result can be generalized to a (symmetric) non-equilibrium configuration for the toy model by the canonical replacement of the equilibrium distribution functions with the non-equilibrium ones: $f^{eq}\rightarrow\f{}{}$. For the phenomenological scenario we obtain in the limit of massless lepton and Higgs for the self-energy contribution (including a factor of 2, because the two components of the lepton doublet can propagate in the self-energy loop for a given transition) \bigformulalign{ \epsilon_i^{S,th} &=-\frac{1}{8\pi}\sum_{j\neq i}\frac{{\Im}\left\{(h^\dagger h)_{ij}^2\right\}}{(h^\dagger h)_{ii}}\frac{M_i M_j}{M_i^2 -M_j^2}\int\frac{\,d\Omega_l}{4\pi}(1-\cos\theta_l)\big\{1-\feq{{\ell}}{E_1} +\feq{\phi}{E_2}\big\}{\, ,} }{CP-violating parameter thermal self-energy} where $E_{1,2}$ are again given by \eqn\eqref{IntermedEnergy}. In the zero temperature limit this reduces to the standard result \bigformulalign{ \epsilon_i^{S,vac}&=-\frac{1}{8\pi}\sum_{j\neq i}\frac{{\Im}\left\{(h^\dagger h)_{ij}^2\right\}}{(h^\dagger h)_{ii}}\frac{M_i M_j}{M_i^2 - M_j^2}{\, .} }{} The complete CP-violating parameter is given by \bigformulalign{ \epsilon_i^{th}&=\epsilon_i^{V,th} +\epsilon_i^{S,th}{\, ,} }{CP-violating parameter thermal complete} where the vertex and self-energy contributions are given by \eqns\eqnref{CP-violating parameter thermal vertex} and \eqnref{CP-violating parameter thermal self-energy} respectively. Therefore the overall dependence on the distribution functions (vertex and self-energy contribution) is given by \bigformula{ 1-\feq{{\ell}}{E_1}+\feq{\phi}{E_2}. }{} In contrast to previous findings \eqn\eqnref{thermal corrections to epsilon phenomenological scenario old}, this does \textit{not} vanish in the limit when the Majorana neutrino decays at rest assuming massless ${\ell}$ and $\phi$. Therefore, it is qualitatively different from the conventional result. The new expression can lead to a significant enhancement of the CP-violating parameter, see \fig\figref{epsilonMediumVsT}. Similar formulas can be derived for processes such as $\phi\rightarrow N_1 {\ell}$ in the standard model (which can become relevant at higher temperatures) or for similar MSSM processes involving sneutrinos and sleptons. The size of the medium corrections depends primarily on the statistics of the particles in the loop, see \fig\figref{epsilonMediumVsT_MSSM}. \section{\label{BoltzmannEq}Boltzmann equations} We can assume in addition that the structure of the {Boltzmann equations} for the phenomenological scenario is analogous to the one given in \cite{Garny:2009rv,Garny:2009qn} with appropriate quantum statistical factors for bosons and fermions respectively and appropriate symmetrization factors. This defines the full set of {Boltzmann equations} including medium corrections to the CP-violating parameter for the phenomenological scenario as derived above. \begin{figure}[!ht] \centering \includegraphics[width=0.7\columnwidth]{figures/epsilonMediumVsT} \caption{Temperature dependence of the CP-violating parameter in the Majorana neutrino decay relative to its vacuum value. Shown are the thermal average $\langle\epsilon_1^{th}\rangle/\epsilon_1^{vac}$ (solid red line) and the values for various momentum modes $\epsilon_1^{th}/\epsilon_1^{vac}$ (dotted red lines) corresponding to $\momqabs=T$, $-1\leq\sin(\delta')\leq+1$. For comparison we also show the conventional results $\langle\epsilon_1^{th,conv}\rangle/\epsilon_1^{vac}$ (dashed black line), where the leading effects cancel as described in the text. Equilibrium distribution functions for bosons and fermions with negligible chemical potentials are assumed. Note that the shown behavior can be modified if thermal masses are included, since the decay $N_1 \rightarrow {\ell}\phi$ (and the conjugate process) becomes kinematically forbidden if the thermal Higgs mass becomes too large. At even higher temperature the process $\phi\rightarrow N_1 {\ell}$ becomes relevant instead \cite{Giudice:2004npb685}.\label{fig:epsilonMediumVsT}} \end{figure} \begin{figure}[!ht] \centering \includegraphics[width=0.7\columnwidth]{figures/epsilonMediumVsT_MSSM} \caption{Medium correction to the CP-violating parameters in the MSSM. The lines correspond to the thermal averages $\langle\epsilon_1^{th}\rangle/\epsilon_1^{vac}$, and the shaded regions illustrate the momentum dependence of $\epsilon_1^{th}/\epsilon_1^{vac}$ for $0.5\leq\momqabs/T\leq 4$ and $\delta'=0$. Note also that in the weighted sum of $\tilde N_1\rightarrow \phi {\tilde \ell}$ and $\tilde N_1\rightarrow \tilde \phi {\ell}$ processes the cancellation of the medium contributions, observed in the earlier publications, does not occur anymore. \label{fig:epsilonMediumVsT_MSSM}} \end{figure} With these modifications, the minimal network of quantum corrected Boltzmann equations for thermal leptogenesis with hierarchical Majorana neutrino masses $M_1 \ll M_2 , M_3$ takes the form (in homogeneous and isotropic Friedman--Robertson--Walker space-time and not writing equations for the Higgs fields $\phi$, $\bar{\phi}$ which are considered to be in thermal equilibrium): \begin{subequations} \label{QuantumBoltzmannEquations} \begin{align} \label{boltzmann equation psi} \lorig [\f{{N}_1}{}](\momkabs)&=\carrayorigkb{{N}_1\leftrightarrow {{\ell}}{\phi}}{\momkabs}{\f{{N}_1}{},\f{{\ell}}{},\f{\phi}{}} +\carrayorigkb{{N}_1\leftrightarrow \bar{{\ell}}\bar{\phi} }{\momkabs}{\f{{N}_1}{},\f{\bar{{\ell}}}{},\f{\bar{\phi}}{}}{\, ,}\\ \label{boltzmann equation b} \lorig [\f{{\ell}}{}](\momkabs)&=\carrayorigkb{{\ell}\phi\leftrightarrow {N}_1}{\momkabs}{\f{{\ell}}{},\f{\phi}{},\f{{N}_1}{}}{\, ,}\\ \label{boltzmann equation bbar} \lorig [\f{\bar{{\ell}}}{}](\momkabs)&=\carrayorigkb{\bar{{\ell}}\bar{\phi} \leftrightarrow {N}_1}{\momkabs}{\f{\bar{{\ell}}}{},\f{\bar{\phi}}{},\f{{N}_1}{}} {\, ,} \end{align} \end{subequations} where the Liouville operator is given by \bigformula{ \lioulong{\f{a}{}}{x}{k} = k^0\left(\frac{\partial }{\partial t}-\abs{\bvec{k}}\hubblerate\frac{\partial }{\partial \abs{\bvec{k}}}\right)\f{a}{}({\abs{\bvec{k}}}){\, .} }{covariant derivative in Friedman Robertson Walker spacetime} If the generated asymmetry is small, as we assume here, then $\f{{\ell}}{}\approx \f{\bar{{\ell}}}{}$ and $\f{\phi}{}\approx \f{\bar{\phi}}{}$. In this case the CP-violating contributions to the right-hand side of \eqn\eqref{boltzmann equation psi} cancel out and we obtain \begin{align} \label{collision term psi-bb and psi-bbar bbar} \carrayorigkb{{N}_1\leftrightarrow {\ell}\phi}{\momkabs}{\f{{N}_1}{},\f{{{\ell}}}{},\f{{\phi}}{}}+ & \carrayorigkb{{N}_1\leftrightarrow \bar{{\ell}}\bar{\phi} }{\momkabs}{\f{{N}_1}{},\f{\bar{{\ell}}}{},\f{\bar{\phi}}{}} \nonumber\\ \simeq{\frac{1}{2}}\int & \lorentzd{{\ell}}{p} \lorentzd{\phi}{q} (2\pi)^4 \delta({k}-{p}-{q})\ampabs{N_1}{{\ell}\phi}^2 (p ,q)\nonumber\\ \times \Big(& \big\{ \qstatferm{\f{{N}_1}{{\momkabs}}}\f{{\ell}}{{\mompabs}}\f{\phi}{{ \momqabs}}-\f{{N}_1}{{\momkabs}}\qstatferm{\f{{\ell}}{{\mompabs}}}\qstat{ \f{\phi}{{\momqabs}}}\big\}\nonumber\\ +&\big\{\qstatferm{\f{{N}_1}{{\momkabs}}}\f{\bar{{\ell}}}{{\mompabs}}\f{ \bar{\phi}}{{\momqabs}} -\f{{N}_1}{{\momkabs}}\qstatferm{\f{\bar{{\ell}}}{{\mompabs}}}\qstat{\f{ \bar{\phi}}{{\momqabs}}}\big\}\Big){\, ,} \end{align} where, as usual, the tree-level amplitude for the Majorana neutrino decay is given by $\ampabs{N_1}{{\ell}\phi}^2 (p,q)=\ampabs{N_1}{\bar{\ell}\bar \phi}^2 (p,q) = 2(h^\dagger h)_{11}\, p\cdotq$. The collision terms for the (inverse) decay of the heavy particle into a ${\ell}\phi$ or a $\bar{{\ell}}\bar{\phi}$ pair explicitly contain the CP-violating parameter $\epsilon_1$ given in \eqn\eqnref{CP-violating parameter thermal complete} but with the equilibrium distributions replaced by non-equilibrium ones $\feq{}{}\rightarrow\f{}{}$: \begin{subequations} \label{collision_term_bbbarbbarb-psi} \begin{align} \label{collision term bb-psi} \carrayorigkb{{\ell}\phi\leftrightarrow {N}_1}{\momkabs}{\f{{\ell}}{},\f{\phi}{},\f{{N}_1}{}}= &{\frac{1}{2}}\int \lorentzd{{\ell}}{p} \lorentzd{{N}_1}{q} (2\pi)^4 \delta({k}+{p}-{q}) \ampabs{N_1}{{\ell}\phi}^2 (k ,p)[1+\epsilon_1(\momqabs)]\nonumber\\ \times & \big\{\qstatferm{\f{{\ell}}{{\momkabs}}}\qstat{\f{\phi}{{\mompabs}}}\f{ {N}_1}{{\momqabs}} -\f{{\ell}}{{\momkabs}}\f{\phi}{{\mompabs}}\qstatferm{\f{{N}_1}{{ \momqabs}}}\big\}{\, ,}\\ \label{collision term bbar bbar-psi} \carrayorigkb{\bar{{\ell}}\bar{\phi} \leftrightarrow {N}_1}{\momkabs}{\f{\bar{{\ell}}}{},\f{\bar{\phi}}{},\f{{N}_1}{}} =&{\frac{1}{2}}\int \lorentzd{\bar{{\ell}}}{p} \lorentzd{{N}_1}{q} (2\pi)^4 \delta({k}+{p}-{q})\ampabs{N_1}{{\ell}\phi}^2 (k ,p)[1-\epsilon_1(\momqabs)]\nonumber\\ \times & \big\{\qstatferm{\f{\bar{{\ell}}}{{\momkabs}}}\qstat{\f{\bar{\phi}}{{\mompabs }}}\f{{N}_1}{{\momqabs}}-\f{\bar{{\ell}}}{{\momkabs}}\f{\bar{\phi}}{{ \mompabs}}\qstatferm{\f{{N}_1}{{\momqabs}}}\big\}{\, .} \end{align} \end{subequations} Note that the network of Boltzmann equations \eqref{QuantumBoltzmannEquations} should be understood in the generalized sense: the transition amplitudes differ from the usual perturbative matrix elements and do not have their symmetry properties as was noted in \cite{Garny:2009rv,Garny:2009qn}. The structure of the collision terms \eqref{collision_term_bbbarbbarb-psi} differs from the conventional one. In particular, we did not include the processes ${\ell}\phi\leftrightarrow \bar{{\ell}}\bar{\phi}$ explicitly, because the collision terms for the processes ${\ell}\phi\leftrightarrow{N}_1$ and $\bar{{\ell}}\bar{\phi}\leftrightarrow{N}_1$ do not suffer from the generation of an asymmetry in equilibrium. To obtain a consistent set of equations in the canonical bottom-up approach we would need to subtract the RIS part of the $S$-matrix element for the processes ${\ell}\phi\leftrightarrow \bar{{\ell}}\bar{\phi}$. Note, however, that it may be necessary to include the collision terms for ${\ell}\phi\leftrightarrow \bar{{\ell}}\bar{\phi}$ (derived in the top-down approach) in quantitative studies, because these can violate {CP} in general. Further scattering processes with top-quarks and gauge-bosons can also give relevant contributions. We note here that this result should be treated with care, because additional new effects could arise when the phenomenological scenario is investigated in the top-down approach. In addition, the applicability of the quasi-particle picture can not be tested in the framework of thermal field theory. In particular the results presented above will only apply in the hierarchical case \cite{Garny:2009qn}. The analysis of the resonant case requires the use of the Kadanoff-Baym formalism, which allows us to take into account the in-medium spectral properties of the mixing fields. \section{\label{Conclusions}Conclusions} Inspired by a discrepancy between conventional results for the thermal corrections to the {CP} asymmetries in thermal leptogenesis and recent new results from non-equilibrium quantum field theory we have reconsidered the calculation of the CP-violating parameters based on thermal quantum field theory. We find that, if causal products are used in the computation of the $n$-point functions, the results of both approaches can be brought into agreement in the framework of a toy model. We conclude that causal $n$-point functions must be used in the derivation of the CP-violating parameter in the phenomenological scenario as well. This leads to new expressions for the thermal corrections to the vertex and self-energy CP-violating parameters. In contrast to the conventional results the thermal corrections do \textit{not} vanish in the limit when the Majorana neutrino decays at rest assuming massless decay products. Therefore, it is qualitatively different from the conventional result and might give significant contributions to the generated baryon asymmetry. In the range from $0.1$ to $10$ of the dimensionless inverse temperature, thermal effects can enhance the CP-violating parameter by up to an order of magnitude. The asymmetry can be computed using the minimal set of Boltzmann equations for leptogenesis in SM+$3\nu_R$ presented here, which are analogous to the equations which have been derived earlier in the framework of the toy model. These take into account decays and inverse decays and include all quantum statistical factors in a way which guarantees that no asymmetry is generated in equilibrium. They can be applied in the case of non-degenerate Majorana neutrino masses. For a detailed phenomenological analysis it will be necessary to take into account further thermal effects such as thermal masses and resummed thermal propagators as well as additional CP-violating processes which exist in phenomenological scenarios. \subsection*{Acknowledgements} \noindent This work was supported by the ``Sonderforschungsbereich'' TR27 and by the ``cluster of excellence Origin and Structure of the Universe''. We would like to thank J-S. Gagnon for useful discussions related to thermal quantum field theory. \gdef\fmfoverset#1#2#3#4{% \parbox{#1}{ {#3} \hspace{-#1} \vspace{-#2} {#4} } } \input{thermal.bbl} \end{document}
\section{Introduction} In condensed matter studies, there are many situations in which molecular dynamics simulation at constant-temperature~\cite{Nose91,Frenkel01,Leimkuhler04a} is needed. For example, this occurs when magnetic systems are modelled in terms of classical spins~\cite{Chen94,Bunker96,Evertz96,Costa97}. Deterministic methods \cite{Nose84,Hoover85,Jellinek88a}, based on non-Hamiltonian dynamics~\cite{Evans90,Tuckerman99,Tuckerman01,Sergi01,Sergi03,Sergi07,Ezra02,Ezra04,Ezra06}, can sample the canonical distribution provided that the motion in the phase space of the relevant degrees of freedom is ergodic \cite{Nose91,Leimkuhler04a}. However, classical spin systems are usually formulated in terms of non-canonical variables \cite{Bulgac90a,Marsden99}, without a kinetic energy expressed through momenta in phase space, so that Nos\'e dynamics cannot be applied directly. To tackle this problem, Bulgac and Kusnezov (BK) introduced a deterministic constant-temperature dynamics~\cite{Bulgac90,Kusnezov90,Kusnezov92} which can be applied to spins. A number of numerical approaches to integration of spin dynamics can be found in the literature~\cite{Frank97,Arponen04,Tsai04,McLachlan06}. However, BK dynamics, as any other deterministic canonical phase space flow, is able to correctly sample the canonical distribution \emph{only} if the motion in phase space is ergodic on the timescale of the simulation. In general, this condition is very difficult to check for statistical systems with many degrees of freedom, while it is known that, despite its simplicity, the one-dimensional harmonic oscillator provides a difficult and important challenge for deterministic thermostatting methods \cite{Hoover85,Hoover01a,Legoll07,Legoll09}. In this paper, we accomplish two goals. First, by reformulating BK dynamics through non-Hamiltonian brackets~\cite{Sergi01,Sergi03} in phase space, we introduce two generalized versions of the BK time evolution which are able to sample the canonical distribution for a stiff harmonic system. Second, using a recently introduced approach based on the geometry of non-Hamiltonian phase space~\cite{Ezra06}, we are able to derive stable and efficient measure-preserving and time-reversible algorithms in a systematic way for all the phase space flows treated here. The BK phase space flow introduces temperature control by means of fictitious coordinates (and their associated momenta in an extended phase space) traditionally called `demons'. Our generalizations of the BK dynamics are obtained by controlling the BK demons themselves by means of additional Nos\'e-type variables \cite{Nose84}. In one case, the BK demons are controlled globally by means of a single additional Nos\'e-Hoover thermostat~\cite{Nose84,Hoover85}. In the following this will be referred to as (BKNH) Bulgac-Kusnezov-Nos\'e-Hoover dynamics. In the second case, each demon is coupled to an independent Nos\'e-Hoover thermostat. This will be called the Bulgac-Kusnezov-Nos\'e-Hoover chain (BKNHC), and corresponds to `massive' NH thermostatting of the demon variables \cite{Martyna92}. The ability to derive numerically stable measure-preserving time-reversible algorithms \cite{Ezra06} for Nos\'e controlled BK dynamics is very encouranging for future applications to thermostatted spin systems. This paper is organized as follows. In Sec.~\ref{sec:nh-bra} we briefly sketch the unified formalism for non-Hamiltonian phase space flows and measure-preserving integration. The BK dynamics is formulated in phase space and a measure-preserving integration algorithm is derived in Sec.~\ref{sec:bk}. The BKNH and BKNH-chain thermostats are treated in Secs.~\ref{sec:cbk} and~\ref{sec:nbkc} respectively. Numerical results for the one-dimensional harmonic oscillator using these thermostats are presented and discussed in Sec.~\ref{sec:num}. Section~\ref{sec:concl} reports our conclusions. In addition we include several Appendices. A useful operator formula is derived in Appendix~\ref{app:app1}, while invariant measures for the BK, BKNH, and BKNHC phase space flows are derived in Appendices \ref{app:BK-stat},~\ref{app:CBK-stat}, and~\ref{app:NBK-stat}, respectively. \newpage \section{Non-Hamiltonian brackets and measure-preserving algorithms} \label{sec:nh-bra} Consider an arbitrary system admitting a time-independent (extended) Hamiltonian expressed in terms of the phase space coordinates $x_i$, $i=1,\ldots, 2N$. In this case, the Hamiltonian can be interpreted as the conserved energy of the system. Upon introducing an antisymmetric tensor field (generalized Poisson tensor \cite{Ramshaw91,Marsden99}) in phase space, $\mbox{\boldmath$\cal B$}(\boldsymbol{x})=-\mbox{\boldmath$\cal B$}^T(\boldsymbol{x})$, one can define non-Hamiltonian brackets~\cite{Sergi01,Sergi03,Sergi07} as \begin{equation} \left\{a,b\right\}=\sum_{i,j=1}^{2n}\frac{\partial a}{\partial x_i} {\cal B}_{ij}\frac{\partial b}{\partial x_j} \;,\label{eq:nh-bra} \end{equation} where $a=a(\boldsymbol{x})$ and $b=b(\boldsymbol{x})$ are two arbitrary phase space functions. The bracket defined in Eq.~(\ref{eq:nh-bra}) is classified as non-Hamiltonian~\cite{Sergi01,Sergi03,Sergi07} since, in general, it does not obey the Jacobi relation, i.e., in general the Jacobiator ${\cal J} \ne 0$, where \cite{Marsden99} \begin{equation} {\cal J}=\left\{a,\left\{b,c\right\}\right\} +\left\{b,\left\{c,a\right\}\right\} +\left\{c,\left\{a,b\right\}\right\}\;, \end{equation} with $c=c(\boldsymbol{x})$ arbitrary phase space function (in addition to the functions $a$ and $b$, previously introduced). If ${\cal J} \neq 0$, the tensor ${\cal B}_{ij}$ is said to define an `almost-Poisson' structure \cite{DaSilva99}. (Such systems have also been called `pseudo-Hamiltonian' \cite{Ramshaw91}.) An energy-conserving and in general non-Hamiltonian phase space flow is then defined by the vector field \begin{equation} \dot{x}_i=\left\{x_i,H\right\}=\sum_{j=1}^{2N}{\cal B}_{ij} \frac{\partial H}{\partial x_j}\;, \label{eq:gen-psf} \end{equation} where conservation of $H(\boldsymbol{x})$ follows directly from the antisymmetry of ${\cal B}_{ij}$. It has previously been shown how equilibrium statistical mechanics can be comprehensively formulated within this framework~\cite{Sergi07}. It is also possible to recast the above formalism and the corresponding statistical mechanics in the language of differential forms~\cite{Ezra02,Ezra04}. If the matrix ${\cal B}$ is invertible (this is true for all the cases considered here), with inverse $\Omega_{ij}$, we can define the 2-form \cite{Schutz80} \begin{equation} \label{eq:2-form} \Omega = \tfrac12 \Omega_{ij} d x^i \wedge d x^j. \end{equation} The dynamics of Eq.~(\ref{eq:gen-psf}) is then Hamiltonian if and only if the form (\ref{eq:2-form}) is \emph{closed}, i.e., has zero exterior derivative, $d \Omega = 0$ \cite{Schutz80}. This condition is independent of the particular system of coordinates used to describe the dynamics. The structure of Eq.~(\ref{eq:gen-psf}) can be taken as the starting point for derivation of efficient time-reversible integration algorithms that also preserve the appropriate measure in phase space~\cite{Ezra06}. Measure-preserving algorithms can be derived upon introducing a splitting of the Hamiltonian \begin{equation} H=\sum_{\alpha=1}^{n_s}H_{\alpha} \label{eq:Hdecomp} \end{equation} which in turn induces a splitting of the Liouville operator associated with the non-Hamiltonian bracket in Eq.~(\ref{eq:nh-bra}), \begin{equation} L_{\alpha}x_i=\left\{x_i,H_{\alpha}\right\}=\sum_{j=1}^{2N}{\cal B}_{ij} \frac{\partial H_{\alpha}}{\partial x_j}\;. \label{eq:Ldecomp} \end{equation} When the phase space flow has a non-zero compressibility \begin{equation} \kappa=\sum_{i,j=1}^{2N}\frac{\partial{\cal B}_{ij}}{\partial x_i} \frac{\partial H}{\partial x_j} \end{equation} the statistical mechanics must be formulated in terms of a modified phase space measure \cite{Tuckerman99,Tuckerman01,Sergi01,Sergi03,Sergi07,Ezra02,Ezra04} \begin{equation} \overline{\omega}=e^{-w(x)}\omega \end{equation} where \begin{equation} \omega=dx^1 \wedge dx^2 \wedge \ldots \wedge dx^{2N} \end{equation} is the standard phase space volume element (volume form \cite{Schutz80}) and the statistical weight $w(x)$ is defined by \begin{equation} \frac{d w }{dt}=\kappa(x)\;. \end{equation} It has been shown that, provided the condition \begin{equation} \label{eq:div} \frac{\partial}{\partial x_j} \;\left[ e^{-w(x)} {\cal B}_{ij}\right] = 0, \;\; i = 1, \ldots 2N \end{equation} is satisfied, then \begin{equation} L_{\alpha}\overline{\omega}=0\quad {\rm for~every}~\alpha\;, \end{equation} so that the volume element $\overline{\omega}$ is invariant under each of the $L_\alpha$~\cite{Ezra06}. The condition (\ref{eq:div}) is satisfied for all the cases considered below, so that, exploiting the decomposition in Eq.~(\ref{eq:Ldecomp}), algorithms derived by means of a symmetric Trotter factorization of the Liouville propagator: \begin{eqnarray} \exp[\tau L]&=& \prod_{\alpha=1}^{n_s-1}\exp\left[\frac{\tau}{2}L_{\alpha}\right] \exp\exp\left[\tau L_{n_x}\right] \nonumber\\ &\times& \prod_{\beta=1}^{n_s-1}\exp\left[\frac{\tau}{2}L_{n_s-\beta}\right] \end{eqnarray} are not only time-reversible but also measure-preserving. \newpage \section{Phase space formulation of the BK thermostat} \label{sec:bk} A phase space formulation of the BK thermostat can be achieved upon introducing the Hamiltonian \begin{subequations} \begin{align} H^{\rm BK} &=\frac{p^2}{2m}+V(q)+\frac{K_1(p_{\zeta})}{m_{\zeta}} +\frac{K_2(p_{\xi})}{m_{\xi}}+k_BT(\zeta+\xi) \\ &=H(q,p)+\frac{K_1(p_{\zeta})}{m_{\zeta}} +\frac{K_2(p_{\xi})}{m_{\xi}}+k_BT(\zeta+\xi)\;, \end{align} \end{subequations} where $(q,p)$ are the physical degrees of freedom (coordinates and momenta), with mass $m$, to be simulated at constant temperature $T$, while $\zeta$ and $\xi$ are the BK `demons', with corresponding inertial parameters $m_{\zeta}$ and $m_{\xi}$, and associated momenta $(p_{\zeta},p_{\xi})$ \cite{Bulgac90,Kusnezov90,Kusnezov92}. $K_1$ and $K_2$ provide the kinetic energy of demon variables, and for the moment are left arbitrary. Upon defining the phase space point as $x=(q,\zeta,\xi,p,p_{\zeta},p_{\xi})=(x_1,x_2,x_3,x_4,x_5,x_6)$, one can introduce an antisymmetric BK tensor field as \begin{equation} \mbox{\boldmath $\cal B$}^{\rm BK} = \left[ \begin{array}{cccccc} 0 & 0 & 0 & 1 & 0 & -G_2 \\ 0 & 0 & 0 & 0 & \frac{\partial G_1}{\partial p} & 0 \\ 0 & 0 & 0 & 0 & 0 & \frac{\partial G_2}{\partial q} \\ -1& 0 & 0 & 0 & -G_1 & 0 \\ 0 & -\frac{\partial G_1}{\partial p}& 0 & G_1 & 0 & 0 \\ G_2 & 0 & -\frac{\partial G_2}{\partial q} & 0 & 0 & 0 \end{array} \right]\;\label{eq:B-BK} \end{equation} where $G_1$ and $G_2$ are functions of system variables $(p, q)$ only. Substituting $\mbox{\boldmath $\cal B$}^{\rm BK}$ and $H^{\rm BK}$ into Eq.~(\ref{eq:gen-psf}), we obtain the energy-conserving equations \begin{subequations} \label{eq:qdotBK} \begin{align} \dot{q}&=\frac{\partial H}{\partial p}-\frac{G_2(q,p)}{m_{\xi}} \frac{\partial K_2}{\partial p_{\xi}} \\ \dot{\zeta}&=\frac{1}{m_{\zeta}}\frac{\partial G_1}{\partial p} \frac{\partial K_1}{\partial p_{\zeta}} \\ \dot{\xi}&=\frac{1}{m_{\xi}}\frac{\partial G_2}{\partial q} \frac{\partial K_2}{\partial p_{\xi}} \\ \dot{p}&=-\frac{\partial H}{\partial q}-\frac{G_1(q,p)}{m_{\zeta}} \frac{\partial K_1}{\partial p_{\zeta}} \\ \dot{p}_{\zeta}&= G_1\frac{\partial H}{\partial p} -k_BT\frac{\partial G_1}{\partial p} \\ \dot{p}_{\xi}&=G_2\frac{\partial H}{\partial q} -k_BT\frac{\partial G_2}{\partial q}\label{eq:pxidotBK}\;. \end{align} \end{subequations} The associated invariant measure for the BK flow is discussed in Appendix \ref{app:BK-stat}. \subsection{Algorithm for BK Dynamics} In order to derive a measure preserving algorithms, the first step, following Eq.~(\ref{eq:Hdecomp}), is to introduce a splitting of $H^{\rm BK}$: \begin{subequations} \begin{align} H_1^{\rm BK}&= V(q) \\ H_2^{\rm BK}&= \frac{p^2}{2m} \\ H_3^{\rm BK}&= k_BT\zeta \\ H_4^{\rm BK}&= k_BT\xi \\ H_5^{\rm BK}&= \frac{K_1(p_{\zeta})}{m_{\zeta}} \\ H_6^{\rm BK}&= \frac{K_2(p_{\xi})}{m_{\xi}}\;. \end{align} \end{subequations} A measure-preserving splitting of the Liouville operator then follows from Eq.~(\ref{eq:Ldecomp}): \begin{subequations} \begin{align} L_1^{\rm BK} & = -\frac{\partial V}{\partial q}\frac{\partial}{\partial p} +G_2\frac{\partial V}{\partial q}\frac{\partial}{\partial p_{\xi}} \\ L_2^{\rm BK} &= \frac{p}{m}\frac{\partial}{\partial q} +G_1\frac{p}{m}\frac{\partial}{\partial p_{\zeta}} \\ L_3^{\rm BK} &= -k_BT\frac{\partial G_1}{\partial p}\frac{\partial}{\partial p_{\zeta}} \\ L_4^{\rm BK} &= -k_BT\frac{\partial G_2}{\partial q}\frac{\partial}{\partial p_{\xi}} \\ L_5^{\rm BK} &= \frac{1}{m_{\zeta}}\frac{\partial G_1}{\partial p} \frac{\partial K_1}{\partial p_{\zeta}}\frac{\partial}{\partial \zeta} -\frac{G_1}{m_{\zeta}} \frac{\partial K_1}{\partial p_{\zeta}}\frac{\partial}{\partial p} \\ L_6^{\rm BK} & = -\frac{G_2}{m_{\xi}}\frac{\partial K_2}{\partial p_{\xi}} \frac{\partial}{\partial q} + \frac{1}{m_{\xi}}\frac{\partial G_2}{\partial q} \frac{\partial K_2}{\partial p_{\xi}} \frac{\partial}{\partial \xi}\;. \end{align} \end{subequations} Upon choosing a symmetric Trotter factorization of the BL Liouville operator based on the decomposition \begin{equation} L^{\rm BK}=\sum_{\alpha=1}^8L_{\alpha}^{\rm BK} \end{equation} a measure-preserving algorithm can be produced in full generality. In practice, a choice of $K_1$, $K_2$, $G_1$, $G_2$ must be made in order obtain explicit formulas. In this paper, we make the following simple choices: \begin{subequations} \label{eq:choiceG1} \begin{align} G_1&= p \label{eq:choicea}\\ G_2&= q\\ K_1&= \frac{p_{\zeta}^2}{2}\\ K_2&= \frac{p_{\xi}^2}{2}\;.\label{eq:choiceb} \end{align} \end{subequations} In terms of Eqs~(\ref{eq:choicea}--\ref{eq:choiceb}), the antisymmetric BK tensor becomes \begin{equation} \tilde{\mbox{\boldmath $\cal B$}}^{\rm BK} = \left[ \begin{array}{cccccc} 0 & 0 & 0 & 1 & 0 & -q \\ 0 & 0 & 0 & 0 & 1 & 0 \\ 0 & 0 & 0 & 0 & 0 & 1 \\ -1& 0 & 0 & 0 & -p & 0 \\ 0 & -1& 0 & p & 0 & 0 \\ q & 0 & -1 & 0 & 0 & 0 \end{array} \right]\;, \end{equation} and the Hamiltonian reads \begin{eqnarray} \tilde{H}^{\rm BK} &=&H(q,p)+\frac{p_{\zeta}^2}{2m_{\zeta}} +\frac{p_{\xi}^2}{2m_{\xi}} +k_BT(\zeta+\xi)\;. \end{eqnarray} The split Liouville operators now simplify as follows: \begin{subequations} \label{eq:bcL1} \begin{align} \tilde{L}_1^{\rm BK} &=-\frac{\partial V}{\partial q}\frac{\partial}{\partial p} +q\frac{\partial V}{\partial q}\frac{\partial}{\partial p_{\xi}} \\ \tilde{L}_2^{\rm BK} &=\frac{p}{m}\frac{\partial}{\partial q} +\frac{p^2}{m}\frac{\partial}{\partial p_{\zeta}} \\ \tilde{L}_3^{\rm BK} &=-k_BT\frac{\partial}{\partial p_{\zeta}} \\ \tilde{L}_4^{\rm BK} &=-k_BT\frac{\partial}{\partial p_{\xi}} \\ \tilde{L}_5^{\rm BK} &= \frac{p_{\zeta}}{m_{\zeta}}\frac{\partial}{\partial \zeta} -\frac{p_{\zeta}}{m_{\zeta}} p\frac{\partial}{\partial p} +\frac{p_{\zeta}^2}{m_{\zeta}}\frac{\partial}{\partial p_{\eta}} \\ \tilde{L}_6^{\rm BK} &= -\frac{p_{\xi}}{m_{\xi}} q\frac{\partial}{\partial q} + \frac{p_{\xi}}{m_{\xi}} \frac{\partial}{\partial \xi} + \frac{p_{\xi}^2}{m_{\xi}} \frac{\partial}{\partial p_{\chi}} \end{align} \end{subequations} For the purposes of defining an efficient integration algorithm, we combine commuting Liouville operators as follows: \begin{subequations} \begin{align} L_A^{\rm BK}&\equiv \tilde{L}_1^{\rm BK}+\tilde{L}_4^{\rm BK} \nonumber \\ & = F(q)\frac{\partial}{\partial p} +F_{p_{\xi}}\frac{\partial}{\partial p_{\xi}} \\ L_B^{\rm BK}&\equiv \tilde{L}_2^{\rm BK}+\tilde{L}_3^{\rm BK} \nonumber\\ &= \frac{p}{m}\frac{\partial}{\partial q} +F_{p_{\zeta}}\frac{\partial}{\partial p_{\zeta}} \\ L_C^{\rm BK}&\equiv \tilde{L}_5^{\rm BK}+\tilde{L}_6^{\rm BK} \nonumber\\ &= -\frac{p_{\zeta}}{m_{\zeta}}p\frac{\partial}{\partial p} -\frac{p_{\xi}}{m_{\xi}}q\frac{\partial}{\partial q} +\frac{p_{\zeta}}{m_{\zeta}}\frac{\partial}{\partial \zeta} +\frac{p_{\xi}}{m_{\xi}}\frac{\partial}{\partial \xi} \end{align} \end{subequations} where \begin{subequations} \begin{align} F(q)&= -\partial V/\partial q \\ F_{p_{\xi}}&= q\frac{\partial V}{\partial q}-k_BT \\ F_{p_{\zeta}}&=\frac{p^2}{m}-k_BT \; . \end{align} \end{subequations} Defining \begin{equation} U^{\rm BK}_{\alpha}(\tau)=\exp\left[\tau \tilde{L}^{\rm BK}_{\alpha}\right]\;, \end{equation} where $\alpha=A,B,C$, one possible reversible measure-preserving integration algorithm for the BK thermostat is then \begin{eqnarray} U(\tau)^{\rm BK}&=& U_B^{\rm BK}\left(\frac{\tau}{4}\right) U_C^{\rm BK}\left(\frac{\tau}{2}\right) U_B^{\rm BK}\left(\frac{\tau}{4}\right) \nonumber\\ &\times& U_A^{\rm BK}\left(\tau\right) \nonumber\\ &\times& U_B^{\rm BK}\left(\frac{\tau}{4}\right) U_C^{\rm BK}\left(\frac{\tau}{2}\right) U_B^{\rm BK}\left(\frac{\tau}{4}\right) \;. \end{eqnarray} Using the so-called direct translation technique \cite{Tuckerman92} we can expand the above symmetric break-up of the Liouville operator into a pseudo-code form, ready to be implemented on the computer: \begin{itemize} \item $ \left. \begin{array}{ccl} q&\to &q+\frac{\tau}{4} \frac{p}{m} \\ p_{\zeta} & \to & p_{\zeta}+ \frac{\tau}{4}F_{p_{\zeta}} \end{array} \right\} : U_B^{\rm BK}\left(\frac{\tau}{4}\right) $ \item $ \left. \begin{array}{ccl} p & \to & p\exp\left[-\frac{\tau}{2} \frac{p_{\zeta}}{m_{\zeta}}\right]\\ q &\to & q\exp\left[-\frac{\tau}{2}\frac{p_{\xi}}{m_{\xi}}\right]\\ \zeta&\to &\zeta+\frac{\tau}{2} \frac{p_{\zeta}}{m_{\zeta}} \\ \xi &\to & \xi +\frac{\tau}{2} \frac{p_{\xi}}{m_{\xi}} \end{array} \right\} : U_C^{\rm BK}\left(\frac{\tau}{2}\right) $ \item $ \left. \begin{array}{ccl} q&\to &q+\frac{\tau}{4} \frac{p}{m} \\ p_{\zeta} & \to & p_{\zeta}+ \frac{\tau}{4}F_{p_{\zeta}} \end{array} \right\} : U_B^{\rm BK}\left(\frac{\tau}{4}\right) $ \item $ \left. \begin{array}{ccl} p&\to &p+\tau F \\ p_{\xi} & \to & p_{\xi}+ \tau F_{p_{\xi}} \end{array} \right\} : U_A^{\rm BK}(\tau) $ \item $ \left. \begin{array}{ccl} q&\to &q+\frac{\tau}{4} \frac{p}{m} \\ \eta &\to & \eta + \frac{\tau}{4} \frac{p_{\eta}}{m_{\eta}} \\ p_{\zeta} & \to & p_{\zeta}+ \frac{\tau}{4}F_{p_{\zeta}} \end{array} \right\} : U_B^{\rm BK}\left(\frac{\tau}{4}\right) $ \item $ \left. \begin{array}{ccl} p & \to & p\exp\left[-\frac{\tau}{2} \frac{p_{\zeta}}{m_{\zeta}}\right]\\ q &\to & q\exp\left[-\frac{\tau}{2}\frac{p_{\xi}}{m_{\xi}}\right]\\ \zeta&\to &\zeta+\frac{\tau}{2} \frac{p_{\zeta}}{m_{\zeta}} \\ \xi &\to & \xi +\frac{\tau}{2} \frac{p_{\xi}}{m_{\xi}} \end{array} \right\} : U_C^{\rm BK}\left(\frac{\tau}{2}\right) $ \item $ \left. \begin{array}{ccl} q&\to &q+\frac{\tau}{4} \frac{p}{m} \\ p_{\zeta} & \to & p_{\zeta}+ \frac{\tau}{4}F_{p_{\zeta}} \end{array} \right\} : U_B^{\rm BK}\left(\frac{\tau}{4}\right) $ \end{itemize} \newpage \section{Bulgac-Kusnezov-Nos\'e-Hoover dynamics} \label{sec:cbk} The BKNH Hamiltonian \begin{equation} H^{\rm BKNH} = H(q,p)+\frac{K_1(p_{\zeta})}{m_{\zeta}} +\frac{K_2(p_{\xi})}{m_{\xi}}+\frac{p_{\eta}^2}{2m_{\eta}} +k_BT(\zeta+\xi) +2k_BT\eta \;\label{eq:H-CBK} \end{equation} is simply the BK Hamiltonian augmented by the Nos\'e variables $(\eta,p_{\eta})$ with mass $m_{\eta}$. With the antisymmetric BKNH tensor \begin{equation} \mbox{\boldmath $\cal B$}^{\rm BKNH} = \left[ \begin{array}{cccccccc} 0 & 0 & 0 & 0 & 1 & 0 & -G_2 & 0 \\ 0 & 0 & 0 & 0 & 0 & \frac{\partial G_1}{\partial p} & 0 & 0 \\ 0 & 0 & 0 & 0 & 0 & 0 & \frac{\partial G_2}{\partial q} & 0 \\ 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \\ -1 & 0 & 0 & 0 & 0 & -G_1 & 0 & 0 \\ 0 & -\frac{\partial G_1}{\partial p}& 0 & 0 & G_1 & 0 & 0 & -p_{\zeta} \\ G_2 & 0 & -\frac{\partial G_2}{\partial q} & 0 & 0 & 0 & 0 & -p_{\xi} \\ 0 & 0 & 0 & -1 & 0 & p_{\zeta} & p_{\xi} & 0 \\ \end{array} \right]\;. \label{eq:B-CBK} \end{equation} we obtain from Eq.~(\ref{eq:gen-psf}) equations of motion for the phase space variables $x=(q,\zeta,\xi,\eta,p,p_{\zeta},p_{\xi},p_{\eta})= (x_1,x_2,x_3,x_4,x_5,x_6,x_7,x_8)$: \begin{subequations} \begin{align} \dot{q}&=\frac{\partial H}{\partial p}-\frac{G_2(q,p)}{m_{\xi}} \frac{\partial K_2}{\partial p_{\xi}} \\ \dot{\zeta}&=\frac{1}{m_{\zeta}}\frac{\partial G_1}{\partial p} \frac{\partial K_1}{\partial p_{\zeta}} \\ \dot{\xi}&=\frac{1}{m_{\xi}}\frac{\partial G_2}{\partial q} \frac{\partial K_2}{\partial p_{\xi}} \\ \dot{\eta}&=\frac{p_{\eta}}{m_{\eta}} \\ \dot{p}&=-\frac{\partial H}{\partial q}-\frac{G_1(q,p)}{m_{\zeta}} \frac{\partial K_1}{\partial p_{\zeta}} \\ \dot{p}_{\zeta}&=G_1\frac{\partial H}{\partial p} -k_BT\frac{\partial G_1}{\partial p}-p_{\zeta}\frac{p_{\eta}}{m_{\eta}} \\ \dot{p}_{\xi}&=G_2\frac{\partial H}{\partial q} -k_BT\frac{\partial G_2}{\partial q}-p_{\xi}\frac{p_{\eta}}{m_{\eta}} \\ \dot{p}_{\eta}&= \frac{p_{\zeta}}{m_{\zeta}} \frac{\partial K_1}{\partial p_{\zeta}} +\frac{p_{\xi}}{m_{\xi}} \frac{\partial K_2}{\partial p_{\xi}} -2k_BT \;. \end{align} \end{subequations} Here, a single Nos\'{e} variable is coupled to both of the BK demons $\zeta$ and $\xi$. The associated invariant measure is discussed in Appendix \ref{app:CBK-stat}. \subsection{Algorithm for BKNH dynamics}\label{sec:CBK-alg} The Hamiltonian can be split as \begin{subequations} \begin{align} H_1^{\rm BKNH}&= V(q) \\ H_2^{\rm BKNH}&= \frac{p^2}{2m} \\ H_3^{\rm BKNH}&= k_BT\zeta \\ H_4^{\rm BKNH}&= k_BT\xi \\ H_5^{\rm BKNH}&= \frac{K_1(p_{\zeta})}{m_{\zeta}} \\ H_6^{\rm BKNH}&= \frac{K_2(p_{\xi})}{m_{\xi}}\\ H_7^{\rm BKNH}&= \frac{p^2_{\eta}}{2m_{\eta}} \\ H_8^{\rm BKNH}&= 2k_BT\eta \end{align} \end{subequations} The measure-preserving splitting~\cite{Ezra06} of the Liouville operator \begin{equation} L_{\alpha}={\cal B}_{ij}^{\rm BKNH} \frac{\partial H^{\rm BKNH}_{\alpha}}{\partial x_j} \frac{\partial}{\partial x_i} \end{equation} yields \begin{subequations} \begin{align} L_1^{\rm BKNH} &=-\frac{\partial V}{\partial q}\frac{\partial}{\partial p} +G_2\frac{\partial V}{\partial q}\frac{\partial}{\partial p_{\xi}} \\ L_2^{\rm BKNH} &=\frac{p}{m}\frac{\partial}{\partial q} +G_1\frac{p}{m}\frac{\partial}{\partial p_{\zeta}} \\ L_3^{\rm BKNH} &= -k_BT\frac{\partial G_1}{\partial p}\frac{\partial}{\partial p_{\zeta}} \\ L_4^{\rm BKNH} &= -k_BT\frac{\partial G_2}{\partial q}\frac{\partial}{\partial p_{\xi}} \\ L_5^{\rm BKNH} &= \frac{1}{m_{\zeta}}\frac{\partial G_1}{\partial p} \frac{\partial K_1}{\partial p_{\zeta}}\frac{\partial}{\partial \zeta} -\frac{G_1}{m_{\zeta}} \frac{\partial K_1}{\partial p_{\zeta}}\frac{\partial}{\partial p} +\frac{p_{\zeta}}{m_{\zeta}} \frac{\partial K_1}{\partial p_{\zeta}}\frac{\partial}{\partial p_{\eta}} \\ L_6^{\rm BKNH} &= -\frac{G_2}{m_{\xi}}\frac{\partial K_2}{\partial p_{\xi}} \frac{\partial}{\partial q} + \frac{1}{m_{\xi}}\frac{\partial G_2}{\partial q} \frac{\partial K_2}{\partial p_{\xi}} \frac{\partial}{\partial \xi} +\frac{p_{\xi}}{m_{\xi}}\frac{\partial K_2}{\partial p_{\xi}} \frac{\partial}{\partial p_{\eta}} \\ L_7^{\rm BKNH} &= \frac{p_{\eta}}{m_{\eta}}\frac{\partial}{\partial \eta} - \frac{p_{\eta}}{m_{\eta}}p_{\zeta} \frac{\partial}{\partial p_{\zeta}} - \frac{p_{\eta}}{m_{\eta}}p_{\xi} \frac{\partial}{\partial p_{\xi}} \\ L_8^{\rm BKNH} &=-2k_BT\frac{\partial}{\partial p_{\eta}} \;. \end{align} \end{subequations} At this stage, we leave the general formulation and adopt the particular choice of $K_1$, $K_2$, $G_1$, and $G_2$ given in Eq.~(\ref{eq:choiceG1}). The antisymmetric BKNH tensor becomes \begin{equation} \tilde{\mbox{\boldmath $\cal B$}}^{\rm BKNH} = \left[ \begin{array}{cccccccc} 0 & 0 & 0 & 0 & 1 & 0 & -q & 0 \\ 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 \\ 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 \\ 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \\ -1 & 0 & 0 & 0 & 0 & -p & 0 & 0 \\ 0 & -1& 0 & 0 & p & 0 & 0 & -p_{\zeta} \\ q & 0 & -1 & 0 & 0 & 0 & 0 & -p_{\xi} \\ 0 & 0 & 0 & -1 & 0 & p_{\zeta} & p_{\xi} & 0 \\ \end{array} \right]\;, \end{equation} and the Hamiltonian simplifies to \begin{equation} \tilde{H}^{\rm BKNH} = H(q,p)+\frac{p_{\zeta}^2}{2m_{\zeta}} +\frac{p_{\xi}^2}{2m_{\xi}}+\frac{p_{\eta}^2}{2m_{\eta}} +k_BT(\zeta+\xi)+2_BT\eta\;. \end{equation} The split Liouville operators are now \begin{subequations} \begin{align} \tilde{L}_1^{\rm BKNH} &=-\frac{\partial V}{\partial q}\frac{\partial}{\partial p} +q\frac{\partial V}{\partial q}\frac{\partial}{\partial p_{\xi}} \label{eq:cL1} \\ \tilde{L}_2^{\rm BKNH} &=\frac{p}{m}\frac{\partial}{\partial q} +\frac{p^2}{m}\frac{\partial}{\partial p_{\zeta}} \\ \tilde{L}_3^{\rm BKNH} &=-k_BT\frac{\partial}{\partial p_{\zeta}} \\ \tilde{L}_4^{\rm BKNH} &=-k_BT\frac{\partial}{\partial p_{\xi}} \\ \tilde{L}_5^{\rm BKNH} &= \frac{p_{\zeta}}{m_{\zeta}} \frac{\partial}{\partial \zeta} -\frac{p_{\zeta}}{m_{\zeta}} p\frac{\partial}{\partial p} +\frac{p_{\zeta}^2}{m_{\zeta}} \frac{\partial}{\partial p_{\eta}} \\ \tilde{L}_6^{\rm BKNH} &= -\frac{p_{\xi}}{m_{\xi}} q\frac{\partial}{\partial q} + \frac{p_{\xi}}{m_{\xi}} \frac{\partial}{\partial \xi} +\frac{p_{\xi}^2}{m_{\xi}} \frac{\partial}{\partial p_{\eta}} \\ \tilde{L}_7^{\rm BKNH} &= \frac{p_{\eta}}{m_{\eta}}\frac{\partial}{\partial \eta} - \frac{p_{\eta}}{m_{\eta}}p_{\zeta} \frac{\partial}{\partial p_{\zeta}} - \frac{p_{\eta}}{m_{\eta}}p_{\xi} \frac{\partial}{\partial p_{\xi}} \\ \tilde{L}_8^{\rm BKNH} &= -2k_BT\frac{\partial}{\partial p_{\eta}} \;. \label{eq:cL10} \end{align} \end{subequations} For the purposes of defining an efficient integration algorithm, we combine commuting Liouville operators as follows: \begin{subequations} \begin{align} L_A^{\rm BKNH}&\equiv \tilde{L}_1^{\rm BKNH}+\tilde{L}_4^{\rm BKNH}+\tilde{L}_7^{\rm BKNH} \nonumber \\ & = F(q)\frac{\partial}{\partial p} +\frac{p_{\eta}}{m_{\eta}}\frac{\partial}{\partial\eta} -\frac{p_{\eta}}{m_{\eta}}p_{\zeta}\frac{\partial}{\partial p_{\zeta}} +\left(-\frac{p_{\chi}}{m_{\chi}}p_{\xi}+F_{p_{\xi}}\right)\frac{\partial}{\partial p_{\xi}} \\ \tilde{L}_B^{\rm BKNH}&\equiv \tilde{L}_2^{\rm BKNH}+\tilde{L}_3^{\rm BKNH} \nonumber\\ &= \frac{p}{m}\frac{\partial}{\partial q} +F_{p_{\zeta}}\frac{\partial}{\partial p_{\zeta}} \\ L_C^{\rm BKNH}&\equiv \tilde{L}_5^{\rm BKNH}+\tilde{L}_6^{\rm BKNH} +\tilde{L}_8^{\rm BKNH} \nonumber\\ &= -\frac{p_{\zeta}}{m_{\zeta}}p\frac{\partial}{\partial p} -\frac{p_{\xi}}{m_{\xi}}q\frac{\partial}{\partial q} +\frac{p_{\zeta}}{m_{\zeta}}\frac{\partial}{\partial \zeta} +\frac{p_{\xi}}{m_{\xi}}\frac{\partial}{\partial \xi} +F_{p_{\eta}}\frac{\partial}{p_{\eta}} \;, \end{align} \end{subequations} where \begin{subequations} \begin{align} F(q)&= -\frac{\partial V}{\partial q} \\ F_{p_{\xi}}&=q\frac{\partial V}{\partial q}-k_BT \\ F_{p_{\zeta}}&=\frac{p^2}{m}-k_BT \\ F_{p_{\eta}}&=\frac{p_{\zeta}^2}{m_{\zeta}}+\frac{p_{\xi}^2}{m_{\xi}} -2k_BT \;. \end{align} \end{subequations} In $L_A$ there appears an operator with the form \begin{equation} \label{eq:operator_CBK} L_i=\left(-\frac{p_k}{m_k}p_i+F_{p_i}\right)\frac{\partial}{\partial p_i} \;, \end{equation} where $(k,i)=(\chi,\xi)$ for $L_A$. The action of the propagator associated with this operator on $p_i$ is derived in Appendix~\ref{app:app1}, and is given by \begin{equation} e^{\tau L_i}p_i = p_ie^{-\tau \frac{p_k}{m_k}}+ \tau F_{p_i}e^{-\tau \frac{p_k}{2m_k}} \left(\tau \frac{p_k}{2m_k}\right)^{-1} \sinh\left[\tau \frac{p_k}{2m_k}\right] \;. \end{equation} The apparently singular function \begin{equation} \left(\tau\frac{p_k}{2m_k}\right)^{-1} \sinh\left[\tau\frac{p_k}{2m_k}\right] \end{equation} is in fact well behaved as $p_k \to 0$, and can be expanded in a Maclaurin series to suitably high order~\cite{Martyna96}. In our implementation we used an eighth order expansion. The propagators for the BKNH dynamics can now be defined as \begin{equation} U_{\alpha}^{\rm BKNH}(\tau)=\exp\left[\tau \tilde{L}_{\alpha}^{\rm BKNH}\right] \;, \end{equation} where $\alpha=A,B,C$. One possible reversible measure-preserving integration algorithm for the BKNH thermostat can then be derived from the following Trotter factorization: \begin{eqnarray} U(\tau)^{\rm BKNH}&=& U_B^{\rm BKNH}\left(\frac{\tau}{4}\right) U_C^{\rm BKNH}\left(\frac{\tau}{2}\right) U_B^{\rm BKNH}\left(\frac{\tau}{4}\right) \nonumber\\ &\times& U_A^{\rm BKNH}\left(\tau\right) \nonumber\\ &\times& U_B^{\rm BKNH}\left(\frac{\tau}{4}\right) U_C^{\rm BKNH}\left(\frac{\tau}{2}\right) U_B^{\rm BKNH}\left(\frac{\tau}{4}\right) \;. \end{eqnarray} The direct translation technique gives the following pseudo-code: \begin{itemize} \item $ \left. \begin{array}{ccl} q&\to &q+\frac{\tau}{4} \frac{p}{m} \\ p_{\zeta} & \to & p_{\zeta}+ \frac{\tau}{4}F_{p_{\zeta}} \end{array} \right\} : U_B^{\rm BKNH}\left(\frac{\tau}{4}\right) $ \item $ \left. \begin{array}{ccl} p & \to & p\exp\left[-\frac{\tau}{2} \frac{p_{\zeta}}{m_{\zeta}}\right]\\ q &\to & q\exp\left[-\frac{\tau}{2}\frac{p_{\xi}}{m_{\xi}}\right]\\ \zeta&\to &\zeta+\frac{\tau}{2} \frac{p_{\zeta}}{m_{\zeta}} \\ \xi &\to & \xi +\frac{\tau}{2} \frac{p_{\xi}}{m_{\xi}} \\ p_{\eta} & \to & p_{\eta} +\frac{\tau}{2}F_{p_{\zeta}} \\ \end{array} \right\} : U_C^{\rm BKNH}\left(\frac{\tau}{2}\right) $ \item $ \left. \begin{array}{ccl} q&\to &q+\frac{\tau}{4} \frac{p}{m} \\ p_{\zeta} & \to & p_{\zeta}+ \frac{\tau}{4}F_{p_{\zeta}} \end{array} \right\} : U_B^{\rm BKNH}\left(\frac{\tau}{4}\right) $ \item $ \left. \begin{array}{ccl} p&\to &p+\tau F(q) \\ p_{\xi} & \to & p_{\xi} + \tau F_{p_{\xi}} \\ \eta &\to & \eta + \tau\frac{p_{\eta}}{m_{\eta}} \\ p_{\zeta} &\to& p_{\zeta}\exp\left[-\tau\frac{p_{\eta}}{m_{\eta}}\right] \end{array} \right\} : U_A^{\rm BKNH}(\tau) $ \item $ \left. \begin{array}{ccl} q&\to &q+\frac{\tau}{4} \frac{p}{m} \\ p_{\zeta} & \to & p_{\zeta}+ \frac{\tau}{4}F_{p_{\zeta}} \end{array} \right\} : U_B^{\rm BKNH}\left(\frac{\tau}{4}\right) $ \item $ \left. \begin{array}{ccl} p & \to & p\exp\left[-\frac{\tau}{2} \frac{p_{\zeta}}{m_{\zeta}}\right]\\ q &\to & q\exp\left[-\frac{\tau}{2}\frac{p_{\xi}}{m_{\xi}}\right]\\ \zeta&\to &\zeta+\frac{\tau}{2} \frac{p_{\zeta}}{m_{\zeta}} \\ \xi &\to & \xi +\frac{\tau}{2} \frac{p_{\xi}}{m_{\xi}} \\ p_{\eta} & \to & p_{\eta} +\frac{\tau}{2}F_{p_{\eta}} \\ \end{array} \right\} : U_C^{\rm BKNH}\left(\frac{\tau}{2}\right) $ \item $ \left. \begin{array}{ccl} q&\to &q+\frac{\tau}{4} \frac{p}{m} \\ p_{\zeta} & \to & p_{\zeta}+ \frac{\tau}{4}F_{p_{\zeta}} \end{array} \right\} : U_B^{\rm BKNH}\left(\frac{\tau}{4}\right) $ \end{itemize} \newpage \section{Bulgac-Kusnezov-Nos\'e-Hoover chain} \label{sec:nbkc} For simplicity, we explicitly treat only the case in which the $p_{\zeta}$ and $p_{\xi}$ demons are each coupled to a standard NH thermostat (length one). It would be straightforward to couple each of the demons to NH chains \cite{Martyna92}, and the general case can be easily inferred from what follows. Define the Hamiltonian \begin{eqnarray} H^{\rm BKNHC} &=&H(q,p)+\frac{K_1(p_{\zeta})}{m_{\zeta}} +\frac{K_2(p_{\xi})}{m_{\xi}}+\frac{p_{\eta}^2}{2m_{\eta}} +\frac{p_{\chi}^2}{2m_{\chi}} +k_BT(\zeta+\xi+\eta+\chi)\;. \label{eq:H-NBK} \end{eqnarray} Upon defining the phase space point $x=(q,\zeta,\xi,\eta,\chi,p,p_{\zeta},p_{\xi},p_{\eta},p_{\chi}) =(x_1,x_2,x_3,x_4,x_5,x_6,x_7,x_8,x_9,x_{10})$ and the antisymmetric BKNHC tensor \begin{equation} \mbox{\boldmath $\cal B$}^{\rm BKNHC} = \left[ \begin{array}{cccccccccc} 0 & 0 & 0 & 0 & 0 & 1 & 0 & -G_2 & 0 & 0\\ 0 & 0 & 0 & 0 & 0 & 0 & \frac{\partial G_1}{\partial p} & 0 & 0 & 0 \\ 0 & 0 & 0 & 0 & 0 & 0 & 0 & \frac{\partial G_2}{\partial q} & 0 & 0 \\ 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 \\ 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \\ -1& 0 & 0 & 0 & 0 & 0 & -G_1 & 0 & 0 & 0 \\ 0 & -\frac{\partial G_1}{\partial p}& 0 & 0 & 0 & G_1 & 0 & 0 & -p_{\zeta} & 0 \\ G_2 & 0 & -\frac{\partial G_2}{\partial q} & 0 & 0 & 0 & 0 & 0 & 0 & -p_{\xi}\\ 0 & 0 & 0 & -1 & 0 & 0 & p_{\zeta} & 0 & 0 & 0\\ 0 & 0 & 0 & 0 & -1 & 0 & 0 & p_{\xi} & 0 & 0\\ \end{array} \right]\;, \label{eq:B-NBK} \end{equation} associated non-Hamiltonian equations of motion are \begin{equation} \dot{x}_i = {\cal B}^{\rm BKNHC}_{ij} \, \frac{\partial H^{{\rm BKNHC}}}{\partial x_j} \end{equation} with $i =1, \ldots, 10$. \iffalse \begin{subequations} \begin{align} \dot{q}&=\frac{\partial H}{\partial p}-\frac{G_2(q,p)}{m_{\xi}} \frac{\partial K_2}{\partial p_{\xi}} \\ \dot{\zeta}&=\frac{1}{m_{\zeta}}\frac{\partial G_1}{\partial p} \frac{\partial K_1}{\partial p_{\zeta}} \\ \dot{\xi}&=\frac{1}{m_{\xi}}\frac{\partial G_2}{\partial q} \frac{\partial K_2}{\partial p_{\xi}} \\ \dot{\eta}&=\frac{p_{\eta}}{m_{\eta}} \\ \dot{\chi}&=\frac{p_{\chi}}{m_{\chi}} \\ \dot{p}&= -\frac{\partial H}{\partial q}-\frac{G_1(q,p)}{m_{\zeta}} \frac{\partial K_1}{\partial p_{\zeta}} \\ \dot{p}_{\zeta}&= G_1\frac{\partial H}{\partial p} -k_BT\frac{\partial G_1}{\partial p}-p_{\zeta}\frac{p_{\eta}}{m_{\eta}} \\ \dot{p}_{\xi}&= G_2\frac{\partial H}{\partial q} -k_BT\frac{\partial G_2}{\partial q}-p_{\xi}\frac{p_{\chi}}{m_{\chi}} \\ \dot{p}_{\eta}&= \frac{p_{\eta}^2}{m_{\zeta}} -k_BT\\ \dot{p}_{\chi}&= \frac{p_{\xi}^2}{m_{\xi}} -k_BT \end{align} \end{subequations} \fi \subsection{Algorithm for BKNHC chain dynamics} Splitting the BKNHC chain Hamiltonian as \begin{subequations} \begin{align} H_1^{\rm BKNHC}&= V(q) \\ H_2^{\rm BKNHC}&= \frac{p^2}{2m} \\ H_3^{\rm BKNHC}&= k_BT\zeta \\ H_4^{\rm BKNHC}&= k_BT\xi \\ H_5^{\rm BKNHC}&= \frac{K_1(p_{\zeta})}{m_{\zeta}} \\ H_6^{\rm BKNHC}&= \frac{K_2(p_{\xi})}{m_{\xi}}\\ H_7^{\rm BKNHC}&= \frac{p^2_{\eta}}{2m_{\eta}} \\ H_8^{\rm BKNHC}&= k_BT\eta\\ H_9^{\rm BKNHC}&= \frac{p^2_{\chi}}{2m_{\chi}} \\ H_{10}^{\rm BKNHC}&= k_BT\chi \;, \end{align} \end{subequations} we obtain the corresponding measure-preserving splitting of the Liouville operator \begin{equation} L_{\alpha}={\cal B}_{ij}^{\rm BKNHC} \frac{\partial H^{\rm BKNHC}_{\alpha}}{\partial x_j} \frac{\partial}{\partial x_i} . \end{equation} \iffalse yields \begin{subequations} \begin{align} L_1^{\rm BKNHC} &=-\frac{\partial V}{\partial q}\frac{\partial}{\partial p} +G_2\frac{\partial V}{\partial q}\frac{\partial}{\partial p_{\xi}} \\ L_2^{\rm BKNHC} &=\frac{p}{m}\frac{\partial}{\partial q} +G_1\frac{p}{m}\frac{\partial}{\partial p_{\zeta}} \\ L_3^{\rm BKNHC} &=-k_BT\frac{\partial G_1}{\partial p}\frac{\partial}{\partial p_{\zeta}} \\ L_4^{\rm BKNHC} &=-k_BT\frac{\partial G_2}{\partial q}\frac{\partial}{\partial p_{\xi}} \\ L_5^{\rm BKNHC} &= \frac{1}{m_{\zeta}}\frac{\partial G_1}{\partial p} \frac{\partial K_1}{\partial p_{\zeta}}\frac{\partial}{\partial \zeta} -\frac{G_1}{m_{\zeta}} \frac{\partial K_1}{\partial p_{\zeta}}\frac{\partial}{\partial p} +\frac{p_{\zeta}}{m_{\zeta}} \frac{\partial K_1}{\partial p_{\zeta}}\frac{\partial}{\partial p_{\eta}} \\ L_6^{\rm BKNHC} &= -\frac{G_2}{m_{\xi}}\frac{\partial K_2}{\partial p_{\xi}} \frac{\partial}{\partial q} + \frac{1}{m_{\xi}}\frac{\partial G_2}{\partial q} \frac{\partial K_2}{\partial p_{\xi}} \frac{\partial}{\partial \xi} + \frac{p_{\xi}}{m_{\xi}}\frac{\partial K_2}{\partial p_{\xi}} \frac{\partial}{\partial p_{\chi}} \\ L_7^{\rm BKNHC} &= \frac{p_{\eta}}{m_{\eta}}\frac{\partial}{\partial \eta} - \frac{p_{\eta}}{m_{\eta}}p_{\zeta} \frac{\partial}{\partial p_{\zeta}} \\ L_8^{\rm BKNHC} &=-k_BT\frac{\partial}{\partial p_{\eta}} \\ L_9^{\rm BKNHC} &= \frac{p_{\chi}}{m_{\chi}}\frac{\partial}{\partial \chi} -\frac{p_{\chi}}{m_{\chi}}p_{\xi} \frac{\partial}{\partial p_{\xi}} \\ L_{10}^{\rm BKNHC} &=-k_BT\frac{\partial}{\partial p_{\chi}} \;. \end{align} \end{subequations} \fi At this stage we go directly to Eqs~(\ref{eq:choiceG1}). The antisymmetric Nos\'e-Hoover-Bulgac-Kusnezov tensor becomes \begin{equation} \tilde{\mbox{\boldmath $\cal B$}}^{\rm BKNHC} = \left[ \begin{array}{cccccccccc} 0 & 0 & 0 & 0 & 0 & 1 & 0 & -q & 0 & 0\\ 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 & 0 \\ 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 \\ 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 \\ 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \\ -1& 0 & 0 & 0 & 0 & 0 & -p & 0 & 0 & 0 \\ 0 & -1& 0 & 0 & 0 & p & 0 & 0 & -p_{\zeta} & 0 \\ q & 0 & -1 & 0 & 0 & 0 & 0 & 0 & 0 & -p_{\xi}\\ 0 & 0 & 0 & -1 & 0 & 0 & p_{\zeta} & 0 & 0 & 0\\ 0 & 0 & 0 & 0 & -1 & 0 & 0 & p_{\xi} & 0 & 0\\ \end{array} \right]\; , \end{equation} the Hamiltonian \begin{eqnarray} \tilde{H}^{\rm BKNHC} &=&H(q,p)+\frac{p_{\zeta}^2}{2m_{\zeta}} +\frac{p_{\xi}^2}{2m_{\xi}}+\frac{p_{\eta}^2}{2m_{\eta}} +\frac{p_{\chi}^2}{2m_{\chi}} +k_BT(\zeta+\xi+\eta+\chi)\; \end{eqnarray} and associated Liouville operators \begin{subequations} \label{eq:L} \begin{align} \tilde{L}_1^{\rm BKNHC} &=-\frac{\partial V}{\partial q}\frac{\partial}{\partial p} +q\frac{\partial V}{\partial q}\frac{\partial}{\partial p_{\xi}} \label{eq:L1} \\ \tilde{L}_2^{\rm BKNHC} &=\frac{p}{m}\frac{\partial}{\partial q} +\frac{p^2}{m}\frac{\partial}{\partial p_{\zeta}} \\ \tilde{L}_3^{\rm BKNHC} &= -k_BT\frac{\partial}{\partial p_{\zeta}} \\ \tilde{L}_4^{\rm BKNHC} &= -k_BT\frac{\partial}{\partial p_{\xi}} \\ \tilde{L}_5^{\rm BKNHC} &= \frac{p_{\zeta}}{m_{\zeta}}\frac{\partial}{\partial \zeta} -\frac{p_{\zeta}}{m_{\zeta}} p\frac{\partial}{\partial p} +\frac{p_{\zeta}^2}{m_{\zeta}}\frac{\partial}{\partial p_{\eta}} \\ \tilde{L}_6^{\rm BKNHC} &= -\frac{p_{\xi}}{m_{\xi}} q\frac{\partial}{\partial q} + \frac{p_{\xi}}{m_{\xi}} \frac{\partial}{\partial \xi} + \frac{p_{\xi}^2}{m_{\xi}} \frac{\partial}{\partial p_{\chi}} \\ \tilde{L}_7^{\rm BKNHC} &= \frac{p_{\eta}}{m_{\eta}}\frac{\partial}{\partial \eta} - \frac{p_{\eta}}{m_{\eta}}p_{\zeta} \frac{\partial}{\partial p_{\zeta}} \\ \tilde{L}_8^{\rm BKNHC} &= -k_BT\frac{\partial}{\partial p_{\eta}} \\ \tilde{L}_{9}^{\rm BKNHC} &= \frac{p_{\chi}}{m_{\chi}}\frac{\partial}{\partial \chi} -\frac{p_{\chi}}{m_{\chi}}p_{\xi} \frac{\partial}{\partial p_{\xi}} \\ \tilde{L}_{10}^{\rm BKNHC} &=-k_BT\frac{\partial}{\partial p_{\chi}} \;. \label{eq:L10} \end{align} \end{subequations} We combine commuting Liouville operators as follows: \begin{subequations} \begin{align} L_A^{\rm BKNHC}&\equiv \tilde{L}_1^{\rm BKNHC}+\tilde{L}_4^{\rm BKNHC} +\tilde{L}_9^{\rm BKNHC} \nonumber \\ & = F(q) \frac{\partial}{\partial p} +\frac{p_{\chi}}{m_{\chi}}\frac{\partial}{\partial\chi} +\left(-\frac{p_{\chi}}{m_{\chi}}p_{\xi}+F_{p_{\xi}}\right)\frac{\partial}{\partial p_{\xi}} \\ L_B^{\rm BKNHC}&\equiv \tilde{L}_2^{\rm BKNHC} +\tilde{L}_3^{\rm BKNHC} +\tilde{L}_7^{\rm BKNHC} \nonumber\\ &= \frac{p}{m}\frac{\partial}{\partial q} +\frac{p_{\eta}}{m_{\eta}}\frac{\partial}{\partial\eta} +\left(-\frac{p_{\eta}}{m_{\eta}}p_{\zeta}+F_{p_{\zeta}}\right)\frac{\partial}{\partial p_{\zeta}} \\ L_C^{\rm BKNHC}&\equiv \tilde{L}_5^{\rm BKNHC}+\tilde{L}_6^{\rm BKNHC} +\tilde{L}_8^{\rm BKNHC}+\tilde{L}_{10}^{\rm BKNHC} \nonumber\\ &= -\frac{p_{\zeta}}{m_{\zeta}}p\frac{\partial}{\partial p} -\frac{p_{\xi}}{m_{\xi}}q\frac{\partial}{\partial q} +\frac{p_{\zeta}}{m_{\zeta}}\frac{\partial}{\partial \zeta} +\frac{p_{\xi}}{m_{\xi}}\frac{\partial}{\partial \xi} +F_{p_{\eta}}\frac{\partial}{p_{\eta}} +F_{p_{\chi}}\frac{\partial}{p_{\chi}} \;, \end{align} \end{subequations} where \begin{subequations} \begin{align} F(q)&=-\frac{\partial V}{\partial q} \\ F_{p_{\xi}}&=q\frac{\partial V}{\partial q}-k_BT \\ F_{p_{\zeta}}&=\frac{p^2}{m}-k_BT \\ F_{p_{\eta}}&=\frac{p_{\zeta}^2}{m_{\zeta}}-k_BT \\ F_{p_{\chi}}&=\frac{p_{\xi}^2}{m_{\xi}}-k_BT \;. \end{align} \end{subequations} Both in $L_A^{\rm BKNHC}$ and $L_B^{\rm BKNHC}$ there appears an operator with the form \begin{equation} L_i=\left(-\frac{p_k}{m_k}p_i+F_{p_i}\right)\frac{\partial}{\partial p_i} \;, \end{equation} where $(k,i)=(\chi,\xi)$ for $L_A$ and $(k,i)=(\eta,\zeta)$ for $L_B$. Again following the derivation in Appendix~\ref{app:app1}, we find \begin{equation} e^{\tau L_i}p_i = p_ie^{-\tau \frac{p_k}{m_k}}+ \tau F_{p_i}e^{-\tau \frac{p_k}{2m_k}} \left(\tau \frac{p_k}{2m_k}\right)^{-1} \sinh\left[\tau \frac{p_k}{2m_k}\right] \;. \end{equation} The function $\left(\tau\frac{p_k}{2m_k}\right)^{-1} \sinh\left[\tau\frac{p_k}{2m_k}\right]$ is treated through an eighth order expansion~\cite{Martyna96}. The propagators \begin{equation} U_{\alpha}^{\rm BKNHC}(\tau)=\exp\left[\tau \tilde{L}_{\alpha}^{\rm BKNHC}\right] \end{equation} with $\alpha=A,B,C$ can now be introduced. One possible reversible measure-preserving integration algorithm for the BKNHC chain thermostat is then \begin{eqnarray} U(\tau)^{\rm BKNHC}&=& U_B^{\rm BKNHC}\left(\frac{\tau}{4}\right) U_C^{\rm BKNHC}\left(\frac{\tau}{2}\right) U_B^{\rm BKNHC}\left(\frac{\tau}{4}\right) \nonumber\\ &\times& U_A^{\rm BKNHC}\left(\tau\right) \nonumber\\ &\times& U_B^{\rm BKNHC}\left(\frac{\tau}{4}\right) U_C^{\rm BKNHC}\left(\frac{\tau}{2}\right) U_B^{\rm BKNHC}\left(\frac{\tau}{4}\right) \;. \end{eqnarray} In pseudo-code form, we have the resulting integration algorithm: \begin{itemize} \item $ \left. \begin{array}{ccl} q&\to &q+\frac{\tau}{4} \frac{p}{m} \\ \eta &\to & \eta + \frac{\tau}{4} \frac{p_{\eta}}{m_{\eta}} \\ p_{\zeta} & \to & p_{\zeta}e^{-\frac{\tau}{4} \frac{p_{\eta}}{m_{\eta}}}+ \frac{\tau}{4}F_{p_{\zeta}}e^{-\frac{\tau}{4} \frac{p_{\eta}}{2m_{\eta}}} \left(\frac{\tau}{4}\frac{p_{\eta}}{2m_{\eta}}\right)^{-1} \sinh\left[\frac{\tau}{4}\frac{p_{\eta}}{2m_{\eta}}\right] \end{array} \right\} : U_B^{\rm BKNHC}\left(\frac{\tau}{4}\right) $ \item $ \left. \begin{array}{ccl} p & \to & p\exp\left[-\frac{\tau}{2} \frac{p_{\zeta}}{m_{\zeta}}\right]\\ q &\to & q\exp\left[-\frac{\tau}{2}\frac{p_{\xi}}{m_{\xi}}\right]\\ \zeta&\to &\zeta+\frac{\tau}{2} \frac{p_{\zeta}}{m_{\zeta}} \\ \xi &\to & \xi +\frac{\tau}{2} \frac{p_{\xi}}{m_{\xi}} \\ p_{\eta} & \to & p_{\eta} +\frac{\tau}{2}F_{p_{\zeta}} \\ p_{\chi} &\to & p_{\chi} +\frac{\tau}{2} F_{p_{\chi}} \end{array} \right\} : U_C^{\rm BKNHC}\left(\frac{\tau}{2}\right) $ \item $ \left. \begin{array}{ccl} q&\to &q+\frac{\tau}{4} \frac{p}{m} \\ \eta &\to & \eta + \frac{\tau}{4} \frac{p_{\eta}}{m_{\eta}} \\ p_{\zeta} & \to & p_{\zeta}e^{-\frac{\tau}{4} \frac{p_{\eta}}{m_{\eta}}}+ \frac{\tau}{4}F_{p_{\zeta}}e^{-\frac{\tau}{4} \frac{p_{\eta}}{2m_{\eta}}} \left(\frac{\tau}{4}\frac{p_{\eta}}{2m_{\eta}}\right)^{-1} \sinh\left[\frac{\tau}{4}\frac{p_{\eta}}{2m_{\eta}}\right] \end{array} \right\} : U_B^{\rm BKNHC}\left(\frac{\tau}{4}\right) $ \item $ \left. \begin{array}{ccl} p&\to &p+\tau F \\ \chi &\to & \chi +\tau \frac{p_{\chi}}{m_{\chi}} \\ p_{\xi} & \to & p_{\xi}e^{-\tau \frac{p_{\chi}}{m_{\chi}}}+ \tau F_{p_{\xi}}e^{-\tau \frac{p_{\chi}}{2m_{\chi}}} \left(\tau\frac{p_{\chi}}{2m_{\chi}}\right)^{-1} \sinh\left[\tau\frac{p_{\chi}}{2m_{\chi}}\right] \end{array} \right\} : U_A^{\rm BKNHC}(\tau) $ \item $ \left. \begin{array}{ccl} q&\to &q+\frac{\tau}{4} \frac{p}{m} \\ \eta &\to & \eta + \frac{\tau}{4} \frac{p_{\eta}}{m_{\eta}} \\ p_{\zeta} & \to & p_{\zeta}e^{-\frac{\tau}{4} \frac{p_{\eta}}{m_{\eta}}}+ \frac{\tau}{4}F_{p_{\zeta}}e^{-\frac{\tau}{4} \frac{p_{\eta}}{2m_{\eta}}} \left(\frac{\tau}{4}\frac{p_{\eta}}{2m_{\eta}}\right)^{-1} \sinh\left[\frac{\tau}{4}\frac{p_{\eta}}{2m_{\eta}}\right] \end{array} \right\} : U_B^{\rm BKNHC}\left(\frac{\tau}{4}\right) $ \item $ \left. \begin{array}{ccl} p & \to & p\exp\left[-\frac{\tau}{2} \frac{p_{\zeta}}{m_{\zeta}}\right]\\ q &\to & q\exp\left[-\frac{\tau}{2}\frac{p_{\xi}}{m_{\xi}}\right]\\ \zeta&\to &\zeta+\frac{\tau}{2} \frac{p_{\zeta}}{m_{\zeta}} \\ \xi &\to & \xi +\frac{\tau}{2} \frac{p_{\xi}}{m_{\xi}} \\ p_{\eta} & \to & p_{\eta} +\frac{\tau}{2}F_{p_{\eta}} \\ p_{\chi} &\to & p_{\chi} +\frac{\tau}{2}F_{p_{\chi}} \end{array} \right\} : U_C^{\rm BKNHC}\left(\frac{\tau}{2}\right) $ \item $ \left. \begin{array}{ccl} q&\to &q+\frac{\tau}{4} \frac{p}{m} \\ \eta &\to & \eta + \frac{\tau}{4} \frac{p_{\eta}}{m_{\eta}} \\ p_{\zeta} & \to & p_{\zeta}e^{-\frac{\tau}{4} \frac{p_{\eta}}{m_{\eta}}}+ \frac{\tau}{4}F_{p_{\zeta}}e^{-\frac{\tau}{4} \frac{p_{\eta}}{2m_{\eta}}} \left(\frac{\tau}{4}\frac{p_{\eta}}{2m_{\eta}}\right)^{-1} \sinh\left[\frac{\tau}{4}\frac{p_{\eta}}{2m_{\eta}}\right] \end{array} \right\} : U_B^{\rm BKNHC}\left(\frac{\tau}{4}\right) $ \end{itemize} \newpage \section{Numerical results} \label{sec:num} In its simplicity, the dynamics of a harmonic mode in one dimension is a paradigmatic example for checking the chaotic (ergodic) properties of constant-temperature phase space flows and the correct sampling of the canonical distribution. It is well known that it is necessary to generalize basic Nos\'e-Hoover dynamics~\cite{Nose84,Hoover85,Nose91} to thermostats such as the Nos\'e-Hoover chain~\cite{Martyna92,Martyna96} in order to produce correct constant-temperature averages for systems such as the harmonic oscillator. Some time ago, BK dynamics was devised to provide a deterministic thermostat for systems such as classical spins \cite{Kusnezov90,Kusnezov92}. To ensure efficient thermostatting, BK found it necessary to introduce several `demons' per thermostatted degree of freedom, where each demon was taken to have a different and in principle complicated coupling to the system degree of freedom \cite{Kusnezov90,Kusnezov92}. In the present work, we keep the form of the system-thermostat coupling as simple as possible, in order to facilitate the formulation of explicit, reversible and measure-preserving integrators \cite{Ezra06}. It is then of interest to investigate the ability of our BK-type thermostats to produce the correct canonical sampling in the case of the harmonic oscillator. Interest in harmonic modes is also justified by the possibility of devising models of condensed matter systems in terms of coupled spins and harmonic modes, as already done in quantum dynamics with so-called spin-boson models \cite{Leggett87}. We therefore investigate the performance of our integration schemes on the simple one-dimensional harmonic oscillator. For the particular calculations reported here, the oscillator angular frequency, all masses and $k_B T$ were taken to be unity. The time step in all cases was $\tau =0.0025$, and all runs were calculated for $10^6$ time steps, starting from the same initial conditions: harmonic oscillator coordinate $q=0.3$, all other phase space variables zero at $t=0$. The measure-preserving algorithms derived here result in stable numerical integration for all the three cases treated: BK, BKNH, and BKNHC chain dynamics. Figure~\ref{fig:fig1} shows the three extended Hamiltonians (normalized by their respective initial time value) versus time. All three Hamiltonians are \emph{numerically} conserved by the corresponding measure-preserving algorithm to very high accuracy (which is maintained in all the three cases). However, the basic BK phase space flow is not capable of producing the correct canonical sampling for a harmonic mode. This can be easily checked since the canonical distribution function of the harmonic oscillator is isotropic in phase space and its radial dependence can be calculated exactly. Details of this way of visualizing the phase space sampling have already been given in~\cite{Sergi01,Sergi03}. Figure~\ref{fig:fig2}, displaying the comparison between the theoretical and the calculated value of the radial probability in phase space, clearly shows that the BK dynamics is not able to produce canonical sampling. A look at the inset of Fig.~\ref{fig:fig2}, showing the phase space distribution for the harmonic mode, also immediately shows that the dynamics is not ergodic. The same analysis has been carried out for BKNH and BKNHC phase space flows, and these are displayed in Fig~\ref{fig:fig3} and Fig~\ref{fig:fig4}, respectively. Within numerical errors, both BKNH and BKNHC thermostats are able to produce the correct canonical distribution for the stiff harmonic modes. Introduction of a single, global Nos\'{e}-type variable in the BKNH thermostat effectively introduces additional coupling between the two demon variables. The effectiveness of the BKNH thermostat is consistent with our findings (results not discussed here) that introduction of explicit coupling between demons in BK thermostat dynamics also leads to efficient thermostatting of the harmonic oscillator. \newpage \section{Conclusions} \label{sec:concl} We have formulated Bulgac-Kusnezov \cite{Kusnezov90,Kusnezov92}, Nos\'e-Hoover controlled Bulgac-Kusnezov, and Bulgac-Kusnezov-Nos\'e-Hoover chain thermostats in phase space by means of non-Hamiltonian brackets \cite{Sergi01,Sergi03}. We have derived time-reversible measure-preserving algorithms \cite{Ezra06} for these three cases and showed that additional control by a single Nos\'e-Hoover thermostat or independent Nos\'e-Hoover thermostats is necessary to produce the correct canonical distribution for a stiff harmonic mode. Measure-preserving dynamics of the kind discussed here is associated with equilibrium simulations (where, for example, there is a single temperature parameter $T$). Stationary phase space distributions associated with non-equilibrium situations are much more complicated than the smooth equilbrium densities analyzed in the present paper \cite{Evans90,Hoover98b,Dorfman99}. Nonequilibrium simulations of heat flow could be carried out by extending the present approach to multimode systems (e.g., a chain of oscillators) coupled to BK-type demons with associated NH thermostats corresponding to two different temperatures \cite{Mundy00,Hoover07,Hoover07a}. The techniques presented here for derivation and implementation of thermostats have been shown to be efficient and versatile. We anticipate that analogous approaches can be usefully applied to systems of classical spins coupled to both harmonic and anharmonic modes. \newpage
\section{Introduction } The study of the exclusive decays in the beauty sector allows one to explore the standard model (SM) and search for new physics effects. The B factory experiments such as BaBar at SLAC, Belle at KEK, LHCb at CERN, and B-TeV at Fermilab make precision tests of the SM and beyond the SM ever more promising. Especially, the $B_s$-meson system becomes a key element in the $B$-physics program of B factories ever since the first evidence for $B_s$ production at the $\Upsilon(5S)$ was found by the CLEO collaboration~\cite{CLEO05,CLEO06}. The D0~\cite{D008} and CDF~\cite{CDF08} Collaborations have made measurements of the charge-parity (CP) violating weak $B_s-\bar{B}_s$ mixing phase $\phi_s$ in $B_s\to J/\psi\phi$ decays. While the SM expectation $\phi^{\rm SM}_s$~\cite{Lenz,Bona} is nearly zero, the measured $\phi_s$ differs from 0 by more than $3\sigma$ (but with a sizable error). This measurement of $\phi_s$ inconsistent with zero (if confirmed) would indicate an evidence of new physics. Recently, the Belle Collaboration also measured the branching ratios of the $B_s\to J/\psi\phi$ and $B_s\to J/\psi\eta$ decays and the preliminary result~\cite{Dr} of the $B_s\to J/\psi\phi$ decay is about 3 times larger than that for the $B_s\to J/\psi\eta$ decay. This ratio agrees with a rough estimate obtained within the naive quark model (neglect octet-singlet mixing), where the $s\bar{s}$ part of the $\eta$ meson wave function is one third in contrast to the fully $s\bar{s}$ content of $\phi$ mesons. With the upcoming chances that a numerous number of $B_s$ mesons will be produced at hadron colliders, one might explore the exclusive rare $B_s$ decays to $(K,\eta,\eta')\ell^+\ell^-$(and $\nu_{\ell}\bar{\nu}_{\ell}$) ($\ell=e,\mu,\tau$) induced by the flavor-changing neutral current (FCNC) transitions $b\to(d,s)$. Since in the SM the rare $B_s$ decays are forbidden at tree level and occur at the lowest order only through one-loop penguin diagrams~\cite{GWS,BM,Misiak,TI,BBL,AMM,KMS,AKS}, the rare $B_s$ decays are well suited to test the SM and detect new physics effects. While the experimental tests of exclusive decays are much easier than those of inclusive ones, the theoretical understanding of exlcusive decays is complicated mainly due to the nonperturbative hadronic form factors entered in the long distance nonperturbative contributions. Therefore, a reliable estimate of the hadronic form factors for the exclusive rare $B_s$ decays is very important to make correct predictions within and beyond the SM. The $\eta-\eta'$ mixing angle may also be extracted from the rare $B_s$ decays to $\eta$ and $\eta'$ final states. In our previous work~\cite{CJK}, we have analyzed the exclusive rare $B\to K\ell^+\ell^-$ decays within the framework of the SM, using our light-front quark model (LFQM) based on the QCD-motivated effective LF Hamiltonian~\cite{CJ1,CJ_PLB1,JC_E}. The experimental values of the branching ratios ${\rm BR}(B\to K\ell^+\ell^-)=(0.75^{+0.25}_{-0.21}\pm 0.09)\times 10^{-6}$ from Belle~\cite{Abe02} and $(0.34\pm 0.07\pm 0.02)\times 10^{-6}$ ($\ell=e,\mu$) from BABAR~\cite{Aubert06} detectors are consistent with our LFQM prediction $0.5\times 10^{-6}$~\cite{CJK} based on the SM. Recently, we also analyzed $B_c$ properties and various exclusive decay modes such as the semileptonic $B_c\to (D,\eta_c,B,B_s)\ell\nu_\ell$ decays~\cite{CJBc}, the rare $B_c\to D_{(s)}\ell^+\ell^-$~\cite{Bcrare}, and the nonleptonic two-body $B_c\to (D_{(s)},\eta_c,B_{(s)})(P,V)$ decays~\cite{CJNRD} (here $P$ and $V$ denote pseudoscalar and vector mesons, respectively). The form factors $f_{\pm}(q^2)$ and $f_T(q^2)$ for the exclusive rare decays~\cite{Bcrare} between two pseudoscalar mesons are obtained in the Drell-Yan-West ($q^+=q^0+q^3=0$) frame~\cite{DYW} (i.e., $q^2=-{\bf q}^2_\perp<0$), which is useful because only the valence contributions are needed unless the zero-mode contribution exists. The covariance (i.e., frame independence) of our model has been checked by performing the LF calculation in the $q^+=0$ frame in parallel with the manifestly covariant calculation using the exactly solvable covariant fermion field theory model in $(3+1)$ dimensions. We also found the zero-mode contribution to the form factor $f_-(q^2)$ and identified~\cite{CJBc} the zero-mode operator that is convoluted with the initial and final state LF wave functions. The purpose of this paper is to extend our our LFQM~\cite{CJK,CJ1,CJ_PLB1,JC_E,CJBc,Bcrare,CJNRD} to calculate the hadronic form factors, decay rates and the longitudinal lepton polarization asymmetries (LPAs) for the exclusive rare $B_s\to (K,\eta, \eta')\ell^+\ell^-$ and $\nu_{\ell}\bar{\nu}_{\ell}$ decays within the SM. The LPA, as another parity-violating observable, is an important asymmetry~\cite{Hew} and could be measured at hadron colliders such as LHCb. In particular, the $\tau$ channel would be more accessible experimentally than $e$- or $\mu$-channels since the LPAs in the SM are known to be proportional to the lepton mass. There are some theoretical approaches to the calculations of the exclusive rare $B_s\to \eta\ell^+\ell^-$~\cite{Skands,Geng,CCD} and $B_s\to \eta'\ell^+\ell^-$~\cite{Geng,CCD} decays, but not the $B_s\to K\ell^+\ell^-$ decay mode as far as we know. The paper is organized as follows. In Sec. 2, the SM operator basis, describing the $b\to (d,s)(\ell^+\ell^-,\nu_{\ell}\bar{\nu}_\ell)$ transitions, is presented. In Sec. 3, we briefly describe the formulation of our LFQM and the procedure of fixing the model parameters using the variational principle for the QCD motivated effective Hamiltonian. We discuss the rare decays between two pseudoscalar mesons using an exactly solvable model based on the covariant Bethe-Salpeter (BS) model of $(3+1)$-dimensional fermion field theory and show the equivalence between the results obtained by the manifestly covariant method and the LF method in the $q^+=0$ frame. We then present the LF covariant forms of the form factors $f_{\pm}(q^2)$ and $f_T(q^2)$ obtained from our LFQM. The $\eta-\eta'$ mixing angle for the $B_s\to\eta^{(\prime)}$ transitions is also discussed in this section. In Sec. 4, our numerical results, i.e. the form factors, decay rates, and the LPAs for the rare $B_s\to (K,\eta,\eta')(\ell^+\ell^-,\nu_{\ell}\bar{\nu}_\ell)$ decays are presented. Summary and discussion of our main results follow in Sec. 5. \section{Effective Hamiltonian} In the SM, the exclusive rare $B_s\to P_q(\ell^+\ell^-, \nu_\ell\bar{\nu}_\ell)$ ($q=d,s$) decays are at the quark level described by the loop $b\to q\;(\ell^+\ell^-, \nu_\ell\bar{\nu}_\ell)$ transitions, and receive contributions from the $Z(\gamma)$-penguin and $W$-box diagrams as shown in Fig.~\ref{fig1}. \begin{figure} \begin{center} \includegraphics[width=3.5in,height=1.5in]{fig1.eps} \end{center} \caption{ Loop diagrams for $B_s\to P_q(\ell^+\ell^-,\nu_\ell\bar{\nu}_\ell)(q=d,s)$ transitions. \label{fig1} } \end{figure} The effective Hamiltonian responsible for the $b\to q\ell^+\ell^-\;(q=d,s)$ decay processes can be represented in terms of the Wilson coefficients, $C_7^{\rm eff},C_9^{\rm eff}$ and $C_{10}$ as~\cite{BM} \bea\label{Hll} {\cal H}^{\ell^+\ell^-}_{\rm eff}&=& \frac{G_F\al_{\rm em}}{2\sqrt{2}\pi}V_{tb}V^*_{tq} \biggl[C^{\rm eff}_9\bar{q}\gamma_\mu (1-\gamma_5) b\bar{\ell}\gamma^\mu\ell + C_{10}\bar{q}\gamma_\mu (1-\gamma_5) b\bar{\ell}\gamma^\mu\gamma_5\ell \nonumber\\ &&-C_7^{\rm eff}\frac{2m_b}{q^2}\bar{q}i\sigma_{\mu\nu}q^\nu(1+\gamma_5) b \bar{\ell}\gamma^\mu\ell \biggr], \eea where $G_F$ is the Fermi constant, $\alpha_{\rm em}$ is the fine structure constant, and $V_{ij}$ are the Cabibbo-Kobayashi-Maskawa (CKM) matrix elements. The relevant Wilson coefficients $C_i$ can be found in Ref.~\cite{BM}. The effective Hamiltonian responsible for the $b\to q\nu_{\ell}\bar{\nu}_{\ell}\;(q=d,s)$ decay processes is given by~\cite{TI,BBL} \bea\label{Hnn} {\cal H}^{\nu_{\ell}\bar{\nu}_{\ell}}_{\rm eff}&=& \frac{G_F\al_{\rm em}}{2\sqrt{2}\pi}V_{tb}V^*_{tq}\frac{X(x_t)}{\sin^2\theta_W} \bar{q}\gamma_\mu (1-\gamma_5) b \bar{\nu}_{\ell}\gamma^\mu(1-\gamma_5)\nu_{\ell}, \eea where $x_t=(m_t/M_W)^2$ and $X(x_t)$ is the top quark loop function~\cite{TI,BBL}, which is given by \be\label{ILF} X(x) = \frac{x}{8}\biggl( \frac{2+x}{x-1} + \frac{3x-6}{(x-1)^2}\ln x\biggr). \ee Besides the short distance (SD) contributions, the main effect on the decay comes from the long distance (LD) contributions due to the $c\bar{c}$ resonance states ($J/\psi,\psi',\cdots$). The effective Wilson coefficient $C^{\rm eff}_9$ taking into account both the SD and LD contributions has the following form~\cite{BM} \be C^{\rm eff}_9(s) = C_9 + Y_{SD}(s) + Y_{LD}(s), \ee where the explicit forms of $Y_{SD}(s)$ and $Y_{LD}(s)$ can be found in~\cite{BM,Faessler}. For the LD contribution $Y_{LD}(s)$, we include two $c\bar{c}$ resonant states $J/\psi(1S)$ and $\psi'(2S)$ and use $\Gamma(J/\psi\to\ell^+\ell^-)=5.26\times 10^{-6}$ GeV, $M_{J/\psi}=3.1$ GeV, $\Gamma_{J/\psi}=87\times 10^{-6}$ GeV for $J/\psi(1S)$ and $\Gamma(\psi'\to\ell^+\ell^-)=2.12\times 10^{-6}$ GeV, $M_{\psi'}=3.69$ GeV, $\Gamma_{\psi'}=277\times 10^{-6}$ GeV for $\psi'(2S)$~\cite{Data08}. The LD contributions to the exclusive $B_s\to P_q\;(q=d,s)$ decays are contained in the meson matrix elements of the bilinear quark currents appearing in ${\cal H}^{\ell^+\ell^-}_{\rm eff}$ and ${\cal H}^{\nu_{\ell}\bar{\nu_{\ell}}}_{\rm eff}$. In the matrix elements of the hadronic currents for $B_s\to P_q$ transitions, the parts containing $\gamma_5$ do not contribute. Considering Lorentz and parity invariances, these matrix elements can be parametrized in terms of hadronic form factors as follows: \be\label{Jmu} J^\mu\equiv\la P_q|\bar{q}\gamma^{\mu} b|B_s\ra= f_{+}(q^{2})P^\mu + f_{-}(q^{2})q^\mu, \ee and \bea\label{JTmu} J^\mu_T&\equiv&\la P_q|\bar{q}i\sigma^{\mu\nu}q_\nu b|B_s\ra = \frac{f_T(q^2)}{M_{B_s}+M_{P_q}}[q^2 P^\mu - (M^2_{B_s}-M^2_{P_q})q^\mu], \eea where $P=P_{B_s}+P_{P_q}$ and $q=P_{B_s}-P_{P_q}$ is the four-momentum transfer to the lepton pair and $4m^{2}_{\ell}\leq q^{2}\leq(M_{B_s}-M_{P_q})^2$. We use the convention $\sigma^{\mu\nu}=(i/2)[\gamma^\mu,\gamma^\nu]$ for the antisymmetric tensor. Sometimes it is useful to express Eq.~(\ref{Jmu}) in terms of $f_+(q^2)$ and $f_0(q^2)$, which are related to the exchange of $1^-$ and $0^+$, respectively, and satisfy the following relations: \be\label{F0} f_+(0)=f_0(0),\; f_0(q^2)=f_+(q^2) + \frac{q^2}{M^2_{B_s}-M^2_{P_q}}f_-(q^2). \ee With the help of the effective Hamiltonian in Eq.~(\ref{Hll}) and Eqs.~(\ref{Jmu}) and~(\ref{JTmu}), the transition amplitude ${\cal M}= \la P_q\ell^+\ell^-|{\cal H}_{\rm eff}|B_s\ra$ for the $B_s\to P_q\ell^+\ell^-$ decay can be written as \bea\label{TranA} {\cal M} &=&\frac{G_F\al_{\rm em}}{2\sqrt{2}\pi}V_{tb}V^*_{tq} \biggl\{\biggl[C^{\rm eff}_9 J_\mu - \frac{2m_b}{q^2}C^{\rm eff}_7 J^T_\mu\biggr] \bar{\ell}\gamma^\mu\ell + C_{10}J_\mu \bar{\ell}\gamma^\mu\gamma_5\ell \biggr\}. \eea The differential decay rate for $B_s\to P_q\ell^+\ell^-$ is given by~\cite{MN,GK} \bea\label{DDR} \frac{d\Gamma_{\ell\ell}}{ds} &=&\frac{M^5_{B_s}G^2_F}{3\cdot2^9\pi^5}\alpha^2_{\rm em} |V_{tb}V^*_{tq}|^2\phi_H^{1/2} \biggl(1-\frac{4t}{s}\biggr)^{1/2} \biggl[\phi_H\biggl(1+\frac{2t}{s}\biggr) {\cal F}_{1} + 12t{\cal F}_{2} \biggr], \eea where \bea\label{DDR2} {\cal F}_{1} &=& \biggl|C^{\rm eff}_9 f_+ - \frac{2\hat{m_b}C^{\rm eff}_7}{1+\sqrt{r}}f_T \biggr|^2 + |C_{10}f_+|^2,\nonumber\\ {\cal F}_{2} &=& |C_{10}|^2 \biggl[(1+r -\frac{s}{2})|f_+|^2 + (1-r)f_+f_- +\frac{s}{2}|f_-|^2\biggr], \nonumber\\ \phi_H &=& (s-1-r)^2-4r, \eea with $s=q^2/M^2_{B_s}$, $t=m^2_\ell/M^2_{B_s}$, $\hat{m_b}=m_b/M_{B_s}$ and $r=M^2_{P_q}/M^2_{B_s}$. Equation~(\ref{DDR}) may be written in terms of ($f_+,f_0,f_T$) instead of ($f_+,f_-,f_T$) as discussed in~\cite{CJK}. Note also from Eqs.~(\ref{DDR}) and~(\ref{DDR2}) that the form factor $f_-(q^2)$ does not contribute in the massless lepton limit. The differential decay rate for $B_s\to P_q\nu_{\ell}\bar{\nu}_{\ell}$ can be easily obtained from the corresponding formula Eq.~(\ref{DDR}) for $B_s\to P_q\ell^+\ell^-$ by the replacement $\hat{m}_\ell\to 0$, $C_7^{\rm eff}\to 0$, and $C^{\rm eff}_9=-C_{10}\to X(x_t)/\sin^2\theta_W$, i.e. \bea\label{DDRn} \frac{d}{ds}\sum_{\ell}\Gamma_{\nu_\ell{\bar\nu}_\ell} &=&3\frac{M^5_{B_s}G^2_F}{3\cdot2^8\pi^5\sin^4\theta_{W}}\alpha^2_{\rm em} |V_{tb}V^*_{tq}|^2\phi_H^{3/2}|X(x_t)|^2|f_+|^2, \eea where the factor of 3 in the numerator corresponds to the sum over the three neutrino flavors. Dividing Eqs.~(\ref{DDR}) and~(\ref{DDRn}) by the total width of the $B_s$ meson, one can obtain the differential branching ratio $d{\rm BR}(B_s\to P_q\ell^+\ell^-)/ds =(d\Gamma(B_s\to P_q\ell^+\ell^-)/\Gamma_{\rm tot})/ds$. As another interesting observable, the LPA, is defined as \be\label{LPA} P_L(s)=\frac{d\Gamma_{h=-1}/ds-d\Gamma_{h=1}/ds} {d\Gamma_{h=-1}/ds +d\Gamma_{h=1}/ds}, \ee where $h=+1(-1)$ denotes right (left) handed $\ell^-$ in the final state. From Eq.~(\ref{DDR}), one obtains for $B_s\to P_q\ell^+\ell^-$ \be\label{LPA_Bc} P_L(s)=\frac{ 2\biggl(1-4\frac{t}{s}\biggr)^{1/2} \phi_H C_{10}f_{+} \biggl[f_+ {\rm Re}C^{\rm eff}_9 - \frac{2\hat{m_b}C^{\rm eff}_7}{1+\sqrt{r}}f_T\biggr] } { \biggl[\phi_H\biggl(1+2\frac{t}{s}\biggr){\cal F}_{1} + 12t{\cal F}_{2} \biggr] }. \ee Because of the experimental difficulties of studying the polarizations of each lepton depending on $s$ and the Wilson coefficients, it would be better to eliminate the dependence of the LPA on $s$, by considering the averaged form over the entire kinematical region. The averaged LPA is defined by \bea\label{LPA_av} \la P_L\ra = \frac{\int^{(1-\sqrt{r})^2}_{4t} P_L\frac{dBR}{ds}ds} {\int^{(1-\sqrt{r})^2}_{4t} \frac{dBR}{ds}ds}. \eea \section{Review of our LFQM} The key idea in our LFQM~\cite{CJ1,CJ_PLB1,CJBc} for the ground state mesons is to treat the radial wave function as a trial function for the variational principle to the QCD-motivated effective Hamiltonian saturating the Fock state expansion by the constituent quark and antiquark. The QCD-motivated effective Hamiltonian for a description of the ground state meson mass spectra is given by \be\label{Ham} H_{q\bar{q}}= H_0 + V_{q\bar{q}}= \sqrt{m^2_q+{\vec k}^2}+\sqrt{m^2_{\bar{q}}+{\vec k}^2}+V_{q\bar{q}}, \ee where \be\label{pot} V_{q\bar{q}}=V_0 + V_{\rm hyp} = a + br^n-\frac{4\al_s}{3r} +\frac{2}{3}\frac{{\bf S}_q\cdot{\bf S}_{\bar{q}}}{m_qm_{\bar{q}}} \nabla^2V_{\rm coul}. \ee In this work, we use the Coulomb plus linear confining (i.e. $n=1$) potential together with the hyperfine interaction $\la{\bf S}_q\cdot{\bf S}_{\bar{q}}\ra=1/4\;(-3/4)$ for the vector (pseudoscalar) meson, which enables us to analyze the meson mass spectra and various wave-function-related observables, such as decay constants, electromagnetic form factors of mesons in a spacelike region, and the weak form factors for the exclusive semileptonic and rare decays of pseudoscalar mesons in the timelike region~\cite{CJK,CJ1,CJ_PLB1,JC_E,CJBc,Bcrare,CJNRD,ChoiRD,Choi08}. The momentum-space LF wave function of the ground state pseudoscalar mesons is given by $\Psi(x_i,{\bf k}_{i\perp},\lam_i) ={\cal R}_{\lam_1\lam_2}(x_i,{\bf k}_{i\perp}) \phi(x_i,{\bf k}_{i\perp})$, where $\phi(x_i,{\bf k}_{i\perp})$ is the radial wave function and ${\cal R}_{\lam_1\lam_2}$ is the covariant spin-orbit wave function. The model wave function is represented by the Lorentz-invariant variables, $x_i=p^+_i/P^+$, ${\bf k}_{i\perp}={\bf p}_{i\perp}-x_i{\bf P}_\perp$ and $\lam_i$, where $P^\mu=(P^+,P^-,{\bf P}_\perp) =(P^0+P^3,(M^2+{\bf P}^2_\perp)/P^+,{\bf P}_\perp)$ is the momentum of the meson $M$, and $p^\mu_i$ and $\lam_i$ are the momenta and the helicities of constituent quarks, respectively. The covariant form of the spin-orbit wave function for pseudoscalar mesons is given by \bea\label{R00_A} {\cal R}_{\lam_1\lam_2} &=&\frac{-\bar{u}_{\lam_1}(p_1)\gamma_5 v_{\lam_2}(p_2)} {\sqrt{2}\sqrt{M^2_0-(m_1-m_2)^2}}, \eea where $M^2_0=\sum_{i=1}^2({\bf k}^2_{i\perp}+m^2_i)/x_i$ is the boost invariant meson mass square obtained from the free energies of the constituents in mesons. For the radial wave function $\phi$, we use the Gaussian wave function: \be\label{rad} \phi(x_i,{\bf k}_{i\perp})=\frac{4\pi^{3/4}}{\beta^{3/2}} \sqrt{\frac{\partial k_z}{\partial x}} {\rm exp}(-{\vec k}^2/2\beta^2), \ee where $\beta$ is the variational parameter. When the longitudinal component $k_z$ is defined by $k_z=(x-1/2)M_0 + (m^2_2-m^2_1)/2M_0$, the Jacobian of the variable transformation $\{x,{\bf k}_\perp\}\to {\vec k}=({\bf k}_\perp, k_z)$ is given by \be\label{jacob} \frac{\partial k_z}{\partial x}=\frac{M_0}{4x_1x_2} \biggl\{ 1- \biggl[\frac{m^2_1-m^2_2}{M^2_0}\biggr]^2\biggr\}. \ee The normalization factor in Eq.~(\ref{rad}) is obtained from the following normalization of the total wave function: \be\label{norm} \int^1_0dx\int\frac{d^2{\bf k}_\perp}{16\pi^3} |\Psi(x,{\bf k}_{i\perp})|^2=1. \ee We apply our variational principle to the QCD-motivated effective Hamiltonian first to evaluate the expectation value of the central Hamiltonian $H_0+V_0$, i.e., $\la\phi|(H_0+V_0)|\phi\ra$, with a trial function $\phi(x_i,{\bf k}_{i\perp})$ that depends on the variational parameter $\beta$. Once the model parameters are fixed by minimizing the expectation value $\la\phi|(H_0+V_0)|\phi\ra$, the mass eigenvalue of each meson is obtained as $M_{q\bar{q}}=\la\phi|(H_0+V_{q\bar{q}})|\phi\ra$. Minimizing energies with respect to $\beta$ and searching for a fit to the observed ground state meson spectra, our central potential $V_0$ obtained from our optimized potential parameters ($a=-0.72$ GeV, $b=0.18$ GeV$^2$, and $\al_s=0.31$)~\cite{CJ1} for the Coulomb plus linear potential was found to be quite comparable with the quark potential model suggested by Scora and Isgur~\cite{SI}, where they obtained $a=-0.81$ GeV, $b=0.18$ GeV$^2$, and $\al_s=0.3\sim 0.6$ for the Coulomb plus linear confining potential. A more detailed procedure for determining the model parameters of light- and heavy-quark sectors can be found in our previous works~\cite{CJ1,CJ_PLB1}. Our model parameters $(m_q,\beta_{q\bar{q}})$ obtained from the linear potential model relevant to this work are summarized in Table~\ref{t1n}. The predictions of the ground state meson mass spectra can be found in~\cite{CJBc}. We should note that our model parameters ($m, \beta$) automatically satisfies the normalization of the total wave function and were fixed by the variational principle to the QCD-motivated effective Hamiltonian. Those parameters in turn automatically satisfies the normalization of the electromagnetic form factors at $q^2=0$ and every other physical observables obtained from our LFQM such as decay constants and electroweak form factors are the predictions. This distinguishes our LFQM from other quark model. \begin{table}[t] \caption{The constituent quark masses[GeV] and the Gaussian parameters $\beta$[GeV] obtained from the linear potential in~\cite{CJ_PLB1}, which are necessary for $B_s\to (K,\eta,\eta')$ decay modes. $q=u$ and $d$.}\label{t1n} \begin{tabular*}{\textwidth}{@{}l*{15}{@{\extracolsep{0pt plus 12pt}}l}} \br $m_q$ & $m_s$ & $m_b$ & $\beta_{qs}$ & $\beta_{ss}$ & $\beta_{sb}$ \\ \mr 0.22 & 0.45 & 5.2 & 0.3886 & 0.4128 & 0.5712 \\ \br \end{tabular*} \end{table} \subsection{Form factors for the rare $B_s\to P$ decays in our LFQM} Most popular phenomenological LFQM uses the Gaussian wave function as the radial wave function due to its predictive power of various physical observables. However, since the LFQM using the Gaussian wave function does not have a counterpart of manifestly covariant model, it is hard to check the covariance of the model. To check the covariance of LFQM, one can start from the manifestly covariant field theory model. For example, using the exactly solvable covariant BS model of (3+1)-dimensional fermion field theory~\cite{BCJ01,MF,Jaus99}, one can perform the LF calculation in parallel with the manifestly covariant calculation and compare the results from the two models. Comparing the LF results and the manifestly covariant results, we were able to derive the LF covariant form factors between two pseudoscalar meson explicitly and to analyze the zero-mode complication. Since the detailed procedure of finding LF covariant transition form factors ($f_+,f_-,f_T$) was already given in our previous works~\cite{CJBc,Bcrare}, we shall briefly describe the essential procedure of obtaining the LF covariant form factors from the exactly solvable covariant BS model of (3+1)-dimensional fermion field theory and show the results of the LF covariant form factors. \begin{figure} \begin{center} \includegraphics[width=4.5in,height=1.5in]{fig2.eps} \end{center} \caption{ The covariant diagram (a) corresponds to the sum of the LF valence diagram (b) defined in $0<k^+<P^+_2$ region and the nonvalence diagram (c) defined in $P^+_2<k^+<P^+_1$ region. The large white and black blobs at the meson-quark vertices in (b) and (c) represent the ordinary LF wave functions and the non-wave-function vertex, respectively. The small black box at the quark-gauge boson vertex indicates the insertion of the relevant Wilson operator. \label{fig2} } \end{figure} The covariant diagram shown in Fig.~\ref{fig2}(a) is in general equivalent to the sum of the LF valence diagram 2(b) and the nonvalence diagram 2(c). The matrix element $J^\mu_{(T)}$ obtained from the covariant diagram of Fig.~\ref{fig2}(a) is given by \be\label{ap:3n} J^\mu_{(T)} = ig_1 g_2 \Lambda^2_1\Lambda^2_2\int\frac{d^4k}{(2\pi)^4} \frac{S^\mu_{(T)}} {N_{\Lambda_1} N_{1} N_{k} N_{2} N_{\Lambda_2}}, \ee where $g_1$ and $g_2$ are the normalization factors which can be fixed by requiring both charge form factors of pseudoscalar mesons to be unity at zero momentum transfer, respectively. To regularize the covariant fermion triangle-loop in (3+1) dimensions, we replace the point gauge-boson vertex $\gamma^\mu$ by a non-local smearing gauge-boson vertex $({\Lambda_1}^2 / N_{\Lambda_1})\gamma^\mu ( {\Lambda_2}^2 / N_{\Lambda_2})$, where $N_{\Lambda_1} =p_1^2-{\Lambda_1}^2+i\ep$ and $N_{\Lambda_2} =p_2^2-{\Lambda_2}^2+i\ep$, and thus the factor $({\Lambda_1}{\Lambda_2})^2$ appears in the normalization factor. $\Lambda_1$ and $\Lambda_2$ play the role of momentum cut-offs similar to the Pauli-Villars regularization~\cite{BCJ01}. Our replacement of $\gamma^\mu$ by the non-local smearing gauge-boson vertex remedies the conceptual difficulty associated with the asymmetry appearing if the fermion loop were regulated by smearing the $q\bar{q}$ bound-state vertex. The rest of the denominators in Eq.~(\ref{ap:3n}), i.e., $N_{1} N_{k} N_{2}$, are coming from the intermediate fermion propagators in the triangle loop diagram and are given by \be\label{ap:4} N_{k} = k^2 - m^2 + i\ep, \; N_{j} = p_j^2 -{m_j}^2 + i\ep \;(j=1,2), \ee where $m_1$, $m$, and $m_2$ are the masses of the constituents carrying the intermediate four-momenta $p_1=P_1 -k$, $k$, and $p_2=P_2 -k$, respectively. Furthermore, the trace terms $S^\mu$ from the vector current and $S^\mu_T$ from the tensor current are given by \bea\label{ap:5} S^\mu &=& {\rm Tr}[\gamma_5(\not\!p_1 + m_1)\gamma^\mu (\not\!p_2 +m_2)\gamma_5(-\not\!k + m)] \nonumber\\ &=&4 \{ p^\mu_1(p_2\cdot k + m_2m) + p^\mu_2 (p_1\cdot k + m_1m) + k^\mu(m_1m_2 - p_1\cdot p_2)\}, \eea and \bea\label{ap:5T} S^\mu_T &=& {\rm Tr}[\gamma_5(\not\!p_1 + m_1)i\sigma^{\mu\nu}q_\nu (\not\!p_2 +m_2)\gamma_5(-\not\!k + m)] \nonumber\\ &=&-4 \{ p^\mu_1 [m(p_2\cdot q) + m_2(k\cdot q)] - p^\mu_2[ m(p_1\cdot q) + m_1(k\cdot q)] \nonumber\\ && + k^\mu [m_1(p_2\cdot q) - m_2(p_1\cdot q)] \}, \eea respectively. By doing the integration over $k^-$ in Eq.~(\ref{ap:3n}), one finds the two LF time-ordered contributions to the residue calculations corresponding to the two poles in $k^-$, the LF valence contribution [Fig.~\ref{fig2}(b)] defined in $0<k^+<P^+_2$ region and the nonvalence contribution [Fig.~\ref{fig2}(c)] defined in $P^+_2<k^+<P^+_1$ region. The nonvalence contribution [Fig.~\ref{fig2}(c)] in the $q^+>0$ frame corresponds to the zero mode (if it exists) in the $q^+\to 0$ limit~\cite{Zero}. Performing the LF calculation of Eq.~(\ref{ap:3n}) in the $q^+=0$ frame in parallel with the manifestly covariant calculation, we use the plus component of the currents to obtain the form factors $f_+(q^2)$ and $f_T(q^2)$. For the form factor $f_-(q^2)$, we use both the plus and perpendicular components of the currents. As we have shown in~\cite{CJBc,Bcrare}, while the form factors $f_+(q^2)$ and $f_T(q^2)$ can be obtained only from the valence contribution in the $q^{+}= 0$ frame without encountering the zero-mode contribution, the form factor $f_-(q^2)$ receives the zero mode. In our recent analysis of semileptonic $B_c$ decays~\cite{CJBc}, we identified the zero-mode operator that is convoluted with the initial and final state LF valence wave functions to generate the zero-mode contribution to the form factor $f_-(q^2)$ in the $q^+=0$ frame. Our method can also be realized effectively by the method presented by Jaus~\cite{Jaus99} using the orientation of the LF plane characterized by the invariant equation $\omega\cdot x=0$, where $\omega$ is an arbitrary light-like four vector. More detailed analysis of the zero-mode operator and the LF covariance of the form factors $f_{\pm}$ and $f_T$ can be found in~\cite{CJBc,Bcrare}. While the manifestly covariant BS model of fermion field theory model is good for the qualitative analysis of the exclusive rare decays, it is still semi-realistic. We thus replace the LF vertex functions in the BS model with the more phenomenological Gaussian radial wave functions in our LFQM since the zero-mode operator is independent from the choice of radial wave function as discussed in~\cite{CJBc}. The LF covariant form factors $f_\pm(q^2)$ and $f_T(q^2)$ for $B_s(q_1\bar{q})\to P(q_2\bar{q})$ transitions obtained from the $q^+=0$ frame are given by (see ~\cite{CJBc, Bcrare} for more detailed derivations) \be\label{fp} f_{+}(q^2) = \int^{1}_{0}dx\int \frac{d^{2}{\bf k}_{\perp}}{16\pi^3} \frac{\phi_{1}(x,{\bf k}_{\perp})}{\sqrt{ {\cal A}_{1}^{2} + {\bf k}^{2}_{\perp}}} \frac{\phi_{2}(x,{\bf k}'_{\perp})}{\sqrt{ {\cal A}_{2}^{2} + {\bf k}^{\prime 2}_{\perp}}} ( {\cal A}_{1}{\cal A}_{2}+{\bf k}_{\perp}\cdot{\bf k'}_{\perp} ), \ee \bea\label{fm} f_-(q^2) &=& \int^1_0 (1-x) dx \int \frac{ d^2{\bf k}_\perp } { 16\pi^3 } \frac{ \phi_1 (x, {\bf k}_\perp) } {\sqrt{ {\cal A}^2_1 + {\bf k}^2_\perp }} \frac{ \phi_2 (x, {\bf k'}_\perp) } {\sqrt{ {\cal A}^2_2 + {\bf k}^{\prime 2}_\perp }} \biggl\{ -x(1-x) M^2_1 \nonumber\\ && - {\bf k}^2_\perp - m_1m + (m_2 - m){\cal A}_1 + 2\frac{q\cdot P}{q^2} \biggl[ {\bf k}^2_\perp + 2\frac{ ( {\bf k}_\perp \cdot {\bf q}_\perp)^2 } {q^2} \biggr] \nonumber\\ && + 2 \frac{ ( {\bf k}_\perp \cdot {\bf q}_\perp)^2 } {q^2} + \frac{ {\bf k}_\perp \cdot {\bf q}_\perp } {q^2} [ M^2_2 - (1-x) (q^2 + q\cdot P) + 2 x M^2_0 \nonumber\\ && - (1 - 2x) M^2_1 - 2(m_1 - m) (m_1 + m_2) ] \biggr\}, \eea \bea\label{fT} f_T(q^2) &=& (M_1 + M_2) \int^1_0 (1-x) dx \int \frac{ d^2{\bf k}_\perp } { 16\pi^3 } \frac{ \phi_1 (x, {\bf k}_\perp) } {\sqrt{ {\cal A}^2_1 + {\bf k}^2_\perp }} \frac{ \phi_2 (x, {\bf k'}_\perp) } {\sqrt{ {\cal A}^2_2 + {\bf k}^{\prime 2}_\perp }} \nonumber\\ &&\times \biggl[ {\cal A}_1 - (m_1 - m_2) \frac{{\bf k}_\perp\cdot{\bf q}_\perp}{ q^2} \biggr], \eea where ${\bf k'}_\perp={\bf k}_\perp + (1-x){\bf q}_\perp$, ${\cal A}_{i}= (1-x) m_{i} + x m$ ($i=1,2$), and $q\cdot P=M^2_1-M^2_2$ with $M_1$ and $M_2$ being the physical masses of the initial and final state mesons, respectively. Our results for the form factors given by Eqs.~(\ref{fp})-(\ref{fT}) are essentially the same as those presented in~\cite{CCH04}. We should note that the LF covariant form factor $f_-(q^2)$ in Eq.~(\ref{fm}) is the sum of the valence contribution $f^{\rm val}_-(q^2)$ and the zero-mode contribution $f^{\rm Z.M.}_-(q^2)$~\cite{CJBc}. Since the form factors $f_{\pm}(q^2)$ and $f_{T}(q^2)$ are defined in the spacelike ($q^2=-{\bf q}^2_\perp <0$) region, we then analytically continue them to the timelike $q^{2}>0$ region by changing ${\bf q}_{\perp}^2$ to $-q^2$ in the form factors. We also compare our analytic solutions with the double pole parametric form given by \be\label{Pole} f_i(q^2)=\frac{f_i(0)}{1-\sigma_1 s+\sigma_2 s^2}, \ee where $\sigma_1$ and $\sigma_2$ are the fitted parameters. \subsection{$\eta-\eta'$ mixing for the $B_s\to \eta^{(\prime)}$ decays } In this subsection, we discuss the $\eta-\eta'$ mixing to obtain the $B_s\to\eta^{(\prime)}$ transition form factors. The octet-singlet mixing angle $\theta$ of $\eta$ and $\eta'$ is known to be in the range of $-10^{\rm o}$ to $-23^{\rm o}$~\cite{Data08}. The physical $\eta$ and $\eta'$ are the mixtures of the flavor $SU(3)$ octet $\eta_8$ and singlet $\eta_0$ states: \be\label{eet} \left( \begin{array}{cc} \eta\\ \eta' \end{array}\,\right) =\left( \begin{array}{cc} \cos\theta\;\; -\sin\theta\\ \sin\theta\;\;\;\;\;\cos\theta \end{array}\,\right)\left( \begin{array}{c} \eta_8\\ \eta_0 \end{array}\,\right), \ee where $\eta_8=(u\bar{u}+d\bar{d}-2s\bar{s})/\sqrt{6}$ and $\eta_0=(u\bar{u}+d\bar{d} + s\bar{s})/\sqrt{3}$. Analogously, in terms of the quark-flavor(QF) basis $\eta_q=(u\bar{u}+d\bar{d})/\sqrt{2}$ and $\eta_s=s\bar{s}$, one obtains~\cite{FKS} \be\label{eea} \left( \begin{array}{cc} \eta\\ \eta' \end{array}\,\right) =\left( \begin{array}{cc} \cos\phi\;\; -\sin\phi\\ \sin\phi\;\;\;\;\;\cos\phi \end{array}\,\right)\left( \begin{array}{c} \eta_q\\ \eta_s \end{array}\,\right). \ee The two schemes are equivalent to each other by $\phi=\theta+ \arctan\sqrt{2}$ when ${\rm SU}_f(3)$ symmetry is perfect. Although it was frequently assumed that the decay constants follow the same pattern of state mixing, the mixing properties of the decay constants will generally be different from those of the meson state since the decay constants only probe the short-distance properties of the valence Fock states while the state mixing refers to the mixing of the overall wave function~\cite{FKS}. Defining $\la P(p)|J^{q(s)}_{\mu5}|0\ra = -if^{q(s)}_P p^\mu$ ($P=\eta,\eta'$) in the QF basis, the four parameters $f^{q}_P$ and $f^{s}_P$ can be expressed in terms of two mixing angles ($\phi_q$ and $\phi_s$) and two decay constants ($f_q$ and $f_s$), i.e.~\cite{FKS}, \be\label{fqfs} \left( \begin{array}{cc} f^q_\eta \;\;\;\;\; f^s_\eta\\ f^q_{\eta'} \;\;\;\;\; f^s_{\eta'} \end{array}\,\right) = \left( \begin{array}{cc} \cos\phi_q\;\; -\sin\phi_s\\ \sin\phi_q\;\;\;\;\;\cos\phi_s \end{array}\,\right)\left( \begin{array}{cc} f_q\;\; 0\\ 0\;\;\;f_s \end{array}\,\right). \ee The difference between the mixing angles $\phi_q-\phi_s$ is due to the Okubo-Zweig-Iizuka(OZI)-violating effects~\cite{OZI} and is found to be small ($\phi_q-\phi_s<5^{\circ}$). The OZI rule implies that the difference between $\phi_q$ and $\phi_s$ vanishes (i.e., $\phi_q=\phi_s=\phi$) to leading order in the $1/N_c$ expansion. Similarly, the four parameters $f^{8}_P$ and $f^{0}_P$ in the octet-singlet basis may be written in terms of two angles ($\theta_8$ and $\theta_0$) and two decay constants ($f_8$ and $f_0$). However, in this case, $\theta_8$ and $\theta_0$ turn out to differ considerably and become equal only in the ${\rm SU}_f(3)$ symmetry limit~\cite{FKS,Leut98}. We shall use the QF basis with the single mixing angle $\phi$ to analyze the $B_s\to\eta^{(\prime)}$ decay modes. In this case, a generic form factor $F$ and the branching ratio for the $B_s\to\eta^{(\prime)}$ are given by \be\label{Fee} F^{B_s\to\eta(\eta')} = -\sin\phi (\cos\phi) F^{B_s\to\eta_s}, \ee with the physical $\eta^{(\prime)}$ mass and \be\label{Gee} {\rm BR}[B_s\to\eta(\eta')\ell^+\ell^-] = \sin^2\phi (\cos^2\phi) {\rm BR}[B_s\to\eta_s\ell^+\ell^-], \ee respectively. Recently, the KLOE Collaboration~\cite{KLOE} extracted the pseudoscalar mixing angle $\phi$ in the QF basis by measuring the ratio ${\rm BR}(\phi\to\eta'\gamma)/{\rm BR}(\phi\to\eta\gamma)$. The measured values are $\phi=(39.7\pm 0.7)^{\circ}$ and $(41.5\pm 0.3_{\rm stat}\pm 0.7_{\rm syst}\pm 0.6_{\rm th})^{\circ}$ with and without the gluonium content for $\eta'$, respectively. However, since the mixing angle for $\eta-\eta'$ is still a controversial issue, we use unspecified value for $\phi$ rather than adopting some specific value. \section{Numerical results} In our numerical calculations for the exclusive rare $B_s\to (K, \eta,\eta')(\nu_{\ell}\bar{\nu_{\ell}}, \ell^+\ell^-)$ decays, we use the model parameters ($m_q,\beta$) for the linear confining potential given in Table~\ref{t1n}. Although our predictions~\cite{CJBc} of ground state heavy meson masses are overall in good agreement with the experimental values, we use the experimental meson masses~\cite{Data08} in the computations of the decay widths to reduce possible theoretical uncertainties. Note that in the numerical calculations we take $(m_c, m_b)=(1.8,5.2)$ GeV in all formulas except in the Wilson coefficient $C^{\rm eff}_9$, where $(m_c, m_{b,\rm pole})=(1.4,4.8)$ GeV have been commonly used~\cite{BM}. For the numerical values of the Wilson coefficients, we use the results given by Ref.~\cite{BM}: \bea\label{WC} C_1 &=&-0.248,\; C_2=1.107, \; C_3=0.011, \nonumber\\ C_4 &=&-0.026, \;C_5=0.007, \; C_6=-0.031, \nonumber\\ C^{\rm eff}_7 &=&-0.313,\; C_9=4.344,\; C_{10}=-4.669, \eea and other input parameters are $|V_{tb}V^*_{ts}|=0.039$, $|V_{tb}V^*_{td}|=0.008$, $\al_{\rm em}^{-1}=129$, $M_W=80.43$ GeV, $m_t=171.3$ GeV, $\sin^2\theta_W=0.2233$, and $\tau_{B^0_s}=1.470$ ps. \begin{figure} \vspace{1cm} \includegraphics[width=3in,height=3in]{fig3a.eps} \hspace{0.5cm} \includegraphics[width=3in,height=3in]{fig3b.eps} \caption{The weak form factors for $B_s\to K$(left panel) and $B_s\to \eta_s$(right panel) decays, respectively.} \label{fig3} \end{figure} \begin{table} \caption{Results for form factors at $q^2=0$ of $B_s\to (K,\eta,\eta')$ transition and parameters $\sigma_i$ defined in Eq.~\protect(\ref{Pole}). The coefficients in $\eta$ and $\eta'$ represent quark mixing angles, i.e. $c_\eta =-\sin\phi$ and $c_{\eta'}=\cos\phi$, respectively.} \begin{tabular*}{\textwidth}{@{}l*{15}{@{\extracolsep{0pt plus 12pt}}l}} \br Mode & $f_+(0)$ & $\sigma_1$ & $\sigma_2$ & $f_-(0)$ & $\sigma_1$ & $\sigma_2$ & $f_T(0)$ & $\sigma_1$ & $\sigma_2$\\ \mr $B_s\to K$ & 0.230 & $-1.650$ & 0.822 & $-0.201$& $-1.638 $ & 0.835 & $-0.228$ & $-1.633 $ & 0.835 \\ $B_s\to\eta$ & $0.291c_\eta$ & $-1.574$ & 0.751 & $-0.231c_\eta$& $-1.582 $ & 0.825 & $-0.280c_\eta$ & $-1.561 $ & 0.782 \\ $B_s\to\eta'$ & $0.291c_{\eta'}$ & $-1.575$ & 0.770 & $-0.225c_{\eta'}$& $-1.570 $ & 0.835 & $-0.300c_{\eta'}$ & $-1.561 $ & 0.802 \\ \br \end{tabular*} \label{t2} \end{table} In Fig.~\ref{fig3}, we show the $q^2$ dependences of the form factors $f_{\pm}(q^2)$ and $f_{T}(q^2)$ for the $B_s\to K$ (left panel) and $B_s\to \eta_s$ with physical masses of $\eta$ and $\eta'$ (right panel), respectively. The form factors at $q^2=0$ and the parameters $\sigma_i$ of the double pole form defined in Eq.~(\ref{Pole}) are listed in Table~\ref{t2}. The form factor $f_+(q^2)$ for the $B_s\to \eta_s$ has the same $q^2$ dependence (apart from the mixing angle $\phi$) for both $\eta$ (solid line) and $\eta'$ (circle) since $f_+$ does not depend on the daughter meson mass as one can see from Eq.~(\ref{fp}). On the other hand, the form factors $f_-$ and $f_T$ between $\eta$ and $\eta'$ are slightly different (apart from the mixing angle $\phi$) since they involve the daughter meson mass as one can see from Eqs.~(\ref{fm}) and ~(\ref{fT}). The form factors at the zero recoil point (i.e., $q^2=q^2_{\rm max})$ correspond to the overlap integral of the initial and final state meson wave functions. The maximum recoil point (i.e., $q^2=0$) corresponds to a final state meson recoiling with the maximum three-momentum $|{\vec P}_f|=(M^2_{B_s}-M^2_f)/2M_{B_s}$ in the rest frame of the $B_s$ meson. As a sensitivity check of our LFQM, we find for the $B_s\to K$ transition that our form factors are changed only about 3$\%$ as the light $d$-quark mass varies about 20$\%$. This indicates that the transition form factors for $b\to d$ decay processes are quite stable on the variation of $d$ quark mass. The form factors for the $B_s\to \eta_s$ have also been computed by Geng and Liu~\cite{Geng} using the similar LFQM but only with the valence contributions in the purely longitudinal $q^+\neq 0$ frame. Although the form factors $f_+$ and $f_T$ at the maximum recoil point obtained from~\cite{Geng} do not receive nonvalence contributions, they receive nonvalence contributions at other nonzero $q^2$ values. The nonvalence contributions to $f_-$ are more serious for the entire $q^2$ range including the $q^2=0$ point. For instance, $f_-$ obtained from~\cite{Geng}(see Fig. 2 in~\cite{Geng}) shows a sharp increasing as $q^2$ near the zero recoil point in contrast to our result. This indicates the nonvalence contribution to $f_-$ is quite large, which in particular overestimate the branching ratio for the $\tau$ dilepton decay mode. Although the form factor $f_-(q^2)$ does not contribute to the branching ratio in the massless lepton ($\ell=e$ or $\mu$) decay, it is important for the heavy $\tau$ decay process. \begin{figure} \vspace{1cm} \includegraphics[width=3in,height=3in]{fig4a.eps} \hspace{0.5cm} \includegraphics[width=3in,height=3in]{fig4b.eps} \caption{Differential branching ratios for $B_s\to K\sum\nu_{\ell}\bar{\nu_{\ell}}$ (left panel) and $B_s\to \eta_s\sum\nu_{\ell}\bar{\nu_{\ell}}$ (right panel) decays, respectively.} \label{fig4} \end{figure} \begin{figure} \vspace{1cm} \includegraphics[width=3in,height=2.3in]{fig5a.eps} \hspace{0.5cm} \includegraphics[width=3in,height=2.3in]{fig5b.eps} \vspace{4ex} \includegraphics[width=3in,height=2.3in]{fig5c.eps} \hspace{0.5cm} \includegraphics[width=3in,height=2.3in]{fig5d.eps} \vspace{4ex} \includegraphics[width=3in,height=2.3in]{fig5e.eps} \hspace{0.5cm} \includegraphics[width=3in,height=2.3in]{fig5f.eps} \caption{Differential branching ratios for $B_s\to K\ell^+\ell^-$ (upper panel) and $B_s\to \eta\ell^+\ell^-$ (middle panel) and $B_s\to \eta'\ell^+\ell^-$ (lower panel) with $\ell=\mu$ and $\tau$, respectively. The solid and dashed lines represent the results without and with the LD contributions, respectively.} \label{fig5} \end{figure} We show our results for the differential branching ratios for $B_s\to (K, \eta_s)\sum\nu_{\ell}\bar{\nu}_{\ell}$ with physical masses of $\eta$ and $\eta'$ in Fig.~\ref{fig4} and $B_s\to (K,\eta,\eta') \ell^+\ell^-$ ($\ell=\mu$ and $\tau$) in Fig.~\ref{fig5}, respectively. For the $B_s\to (K,\eta,\eta')\ell^+\ell^-$ transitions in Fig.~\ref{fig5}, the solid (dashed) lines represent the results without (with) the LD contribution to $C^{\rm eff}_9$. Since the $B_s\to K$ is induced by $b\to d$ transition compared to the $B_s\to\eta_s$ induced by $b\to s$ at the quark level, the branching ratios for the final $K$ meson are order of magnitude smaller than the corresponding branching ratios for the final $\eta$ meson. For the $B_s\to (K,\eta,\eta') \ell^+\ell^-$ decays (see Fig.~\ref{fig5}), the LD contributions (dashed lines) clearly overwhelm the nonresonant branching ratios near $J/\psi(1S)$ and $\psi'(2S)$ peaks, however, suitable $\ell^+\ell^-$ invariant mass cuts can separate the LD contribution from the SD one away from these peaks. This divides the spectrum into two distinct regions~\cite{Hew,AGM}: (i) low-dilepton mass, $4m^2_\ell\leq q^2\leq M^2_{J/\psi}-\delta$, and (ii) high-dilepton mass, $M^2_{\psi'}+\delta\leq q^2\leq q^2_{\rm max}$, where $\delta$ is to be matched to an experimental cut. Our predictions for the nonresonant branching ratios are summarized in Table~\ref{t3} with general form of the mixing angle $\phi$ in the QF basis. Our results are also compared with other theoretical predictions such as the LF and constituent QM~\cite{Geng} and the QCD sum rules (SR)~\cite{CCD} within the SM. Since the amplitude $B_s\to (K,\eta,\eta')\ell^+\ell^-$ is regular at $q^2=0$, the transitions $B_s\to (K,\eta, \eta')e^+e^-$ and $B_s\to (K,\eta,\eta')\mu^+\mu^-$ have almost the same decay rates, i.e. insensitive to the mass of the light lepton. Our predictions of branching ratios are close to the QCD SR results~\cite{CCD} but a bit smaller than the LFQM results~\cite{Geng}. But the results from~\cite{Geng} could be lowered if the nonvalence contributions are properly taken into account. For the mixing angle $\theta=-20^{\circ}$ ($-10^{\circ}$) in the octet-singlet basis, which corresponds to $\phi=34.74^{\circ}$ ($44.74^{\circ}$) in the QF basis, we obtain ${\rm BR}(B_s\to \eta\sum\nu_{\ell}\bar{\nu}_{\ell})=1.1\; (1.7)\times 10^{-6}$, ${\rm BR}(B_s\to \eta\mu^+\mu^-)=1.5 \;(2.4)\times 10^{-7}$, ${\rm BR}(B_s\to \eta\tau^+\tau^-)=3.8 \;(5.8)\times 10^{-8}$, ${\rm BR}(B_s\to \eta'\sum\nu_{\ell}\bar{\nu}_{\ell})=1.8\; (1.3)\times 10^{-6}$, ${\rm BR}(B_s\to \eta'\mu^+\mu^-)=2.4 \;(1.8)\times 10^{-7}$, and ${\rm BR}(B_s\to \eta'\tau^+\tau^-)=3.4 \;(2.6)\times 10^{-8}$, respectively. It is also worth comparing the branching ratios between $B_s\to K$ and $B_s\to\eta$, which may be written as \be \frac{{\rm BR}(B_s\to K\mu^+\mu^-)}{{\rm BR}(B_s\to \eta\mu^+\mu^-)} = \frac{1}{\sin^2\phi}\biggl|\frac{V_{td}}{V_{ts}}\biggr|^2 (1 - \Delta_{SU(3)}), \ee where the $SU(3)$ correction term $\Delta_{SU(3)}$ is estimated about $0.3$ in our model calculation. Such a kind of relation may be further scrutinized by considering an additional correction term neglected in the effective Hamiltonian as discussed in~\cite{CCD}. The branching ratios with the LD contributions for $B_s\to (K,\eta,\eta')\ell^+\ell^-$ $(\ell=\mu,\tau)$ are also presented in Table~\ref{t4} for low- and high-dilepton mass regions of $q^2$. \begin{table} \caption{Nonresonant branching ratios (in units of $10^{-7}$) for $B_s\to (K,\eta,\eta')\sum\nu_{\ell}\bar{\nu}_{\ell}$ and $B_s\to (\eta,\eta')\ell^+\ell^-$ transitions compared with other theoretical model predictions within the SM.} \begin{tabular*}{\textwidth}{@{}l*{15}{@{\extracolsep{0pt plus 12pt}}l}} \br Mode & This work & ~\protect\cite{Geng} & ~\protect\cite{CCD} \\ \mr $B_s\to K\sum\nu_{\ell}\bar{\nu}_{\ell}$ & 1.01 & & \\ $B_s\to \eta\sum\nu_{\ell}\bar{\nu}_{\ell}$ & $35.1\sin^2\phi$ & $58.3\sin^2\phi$(LFQM) & $(21.6\pm 4.6)\sin^2\phi$(set A) \\ & & $54.1\sin^2\phi$(CQM) & $(50.1\pm 15.9)\sin^2\phi$(set B) \\ $B_s\to \eta'\sum\nu_{\ell}\bar{\nu}_{\ell}$ & $26.2\cos^2\phi$ & $42.1\cos^2\phi$(LFQM) & $(16.0\pm 3.6)\cos^2\phi$ (set A) \\ & & $39.7\cos^2\phi$(CQM) & $(33.9\pm 8.9)\cos^2\phi$ (set B) \\ $B_s\to K\mu^+\mu^-$ & 0.14 & & \\ $B_s\to \eta\mu^+\mu^-$ & $4.75\sin^2\phi$ & $8.53\sin^2\phi$(LFQM) & $(2.73\pm 0.68)\sin^2\phi$(set A) \\ & & $7.78\sin^2\phi$(CQM) & $(5.92\pm 1.59)\sin^2\phi$ (set B) \\ $B_s\to \eta'\mu^+\mu^-$ & $3.53\cos^2\phi$ & $6.06\cos^2\phi$(LFQM) & $(1.96\pm 0.53)\cos^2\phi$(set A) \\ & & $5.69\cos^2\phi$(CQM) & $(3.92\pm 1.07)\cos^2\phi$ (set B) \\ $B_s\to K\tau^+\tau^-$ & 0.03 & & \\ $B_s\to \eta\tau^+\tau^-$ & $1.17\sin^2\phi$ & $1.67\sin^2\phi$(LFQM) & $(0.68\pm 0.11)\sin^2\phi$(set A) \\ & & $1.67\sin^2\phi$(CQM) & $(1.82\pm 0.34)\sin^2\phi$ (set B) \\ $B_s\to \eta'\tau^+\tau^-$ & $0.51\cos^2\phi$ & $0.83\cos^2\phi$(LFQM) & $(0.28\pm 0.05)\cos^2\phi$(set A) \\ & & $0.72\cos^2\phi$(CQM) & $(0.69\pm 0.13)\cos^2\phi$ (set B) \\ \br \end{tabular*} \label{t3} \end{table} \begin{table} \caption{Branching ratios with the LD contributions for $B_c\to (D,D_s)\ell^+\ell^-$ for low and high dilepton mass regions of $q^2$ [GeV$^2$] obtained from the linear (HO) potential parameters.} \begin{tabular*}{\textwidth}{@{}l*{15}{@{\extracolsep{0pt plus 12pt}}l}} \br Mode & $4m^2_\ell\leq q^2\leq 8.5$& $14.5\leq q^2\leq q^2_{\rm max}$ \\ \mr $B_s\to K\mu^+\mu^-$ & $7.72\;(6.63)\times10^{-9}$ & $2.27 \;(2.62)\times10^{-9}$ \\ $B_s\to K\tau^+\tau^-$ & & $2.43 \;(2.66)\times10^{-9}$ \\ $B_s\to \eta\mu^+\mu^-$ & $2.86 \;(2.44)\sin^2\phi\times10^{-7}$ & $7.00 \;(8.10)\sin^2\phi\times10^{-8}$ \\ $B_s\to \eta\tau^+\tau^-$ & & $9.13 \;(9.85)\sin^2\phi\times10^{-8}$ \\ $B_s\to \eta'\mu^+\mu^-$ & $2.54 \;(2.17)\cos^2\phi\times10^{-7}$ & $2.54 \;(2.13)\cos^2\phi\times10^{-8}$\\ $B_s\to \eta'\tau^+\tau^-$ & & $3.42 \;(3.67)\cos^2\phi\times10^{-8}$ \\ \br \end{tabular*} \label{t4} \end{table} In Fig.~\ref{fig6}, we show the LPAs for $B\to (K,\eta,\eta')\ell^+\ell^-$ ($\ell=\mu,\tau$) as a function of $s$. In both $\mu$ and $\tau$ dilepton decays, the LPAs become zero at the end point regions of $s$. However, we note that if $m_\ell =0$, the LPAs are not zero at the end points. As in the case of the $B\to K\mu^+\mu^-$~\cite{CJK,GK,MN,HQ} and $B_c\to D_{(s)}\mu^+\mu^-$~\cite{Bcrare} decays where $P_L\simeq -1$ away from the end point regions, the LPAs away from the end point regions are also close to $-1$ for the $B_s\to (K,\eta,\eta')\mu^+\mu^-$ transitions. In fact, the $P_L$ for the muon decay is insensitive to the form factors, e.g. our $P_L$ away from the end point regions is well approximated by~\cite{HQ} \be\label{PLmu} P_L\simeq 2\frac{C_{10}{\rm Re}C^{\rm eff}_9}{|C^{\rm eff}_9|^2 + |C_{10}|^2}\simeq -1, \ee in the limit of $C^{\rm eff}_7\to 0$ from Eq.~(\ref{LPA_Bc}). It also shows that the $P_L$ for the $\mu$ dilepton channel is insensitive to the little variation of $C^{\rm eff}_7$ as expected. On the other hand, the LPA for the $\tau$ dilepton channel is sensitive to the form factors. Similar observation has also been made in our recent work for $B_c\to (D,D_s)\ell^+\ell^-$ decays~\cite{Bcrare}. The averaged values $\la P^{K,\eta^{(\prime)}}_L\ra_\ell$ of the LPAs for $B_s\to (K,\eta^{(\prime)})\ell^+\ell^-$ without the LD contributions are $\la P^K_L\ra_\mu=\la P^\eta_L\ra_\mu=\la P^{\eta'}_L\ra_\mu=-0.98$, $\la P^K_L\ra_\tau=-0.24$, $\la P^\eta_L\ra_\tau=-0.20$ and $\la P^{\eta'}_L\ra_\tau=-0.14$, respectively. \begin{figure} \vspace{1cm} \includegraphics[width=3in,height=2.3in]{fig6a.eps} \hspace{0.5cm} \includegraphics[width=3in,height=2.3in]{fig6b.eps} \vspace{4ex} \includegraphics[width=3in,height=2.3in]{fig6c.eps} \hspace{0.5cm} \includegraphics[width=3in,height=2.3in]{fig6d.eps} \vspace{4ex} \includegraphics[width=3in,height=2.3in]{fig6e.eps} \hspace{0.5cm} \includegraphics[width=3in,height=2.3in]{fig6f.eps} \caption{Longitudinal lepton polarization asymmetries for $B_s\to K\ell^+\ell^-$(upper panel), $B_s\to \eta\ell^+\ell^-$(middle panel) and $B_s\to \eta'\ell^+\ell^-$(lower panel). The same line codes are used as in Fig.~\ref{fig5}.} \label{fig6} \end{figure} \section{Summary and Discussion} In this work, we investigated the exclusive rare semileptonic $B_s\to (K,\eta,\eta')(\nu_{\ell}\bar{\nu_{\ell}},\ell^+\ell^-)$ ($\ell=e,\mu,\tau$) decays within the SM, using our LFQM constrained by the variational principle for the QCD motivated effective Hamiltonian with the linear plus Coulomb interaction~\cite{CJ1,CJ_PLB1}. Our model parameters obtained from the variational principle uniquely determine the physical quantities related to the above processes. This approach can establish the broader applicability of our LFQM to the wider range of hadronic phenomena. The weak form factors $f_{\pm}(q^2)$ and $f_T(q^2)$ for the $B_s\to (K,\eta,\eta')$ decays are obtained in the $q^+=0$ frame ($q^2=-{\bf q}^2_\perp<0$) and then analytically continued to the timelike region by changing ${\bf q}^2_\perp$ to $-q^2$ in the form factors. The covariance (i.e., frame independence) of our model has been checked by performing the LF calculation in parallel with the manifestly covariant calculation using the exactly solvable covariant fermion field theory model in $(3+1)$-dimensions. While the form factors $f_+(q^2)$ and $f_T(q^2)$ are immune to the zero modes, the form factor $f_-(q^2)$ is not free from the zero mode. Using the solutions of the weak form factors obtained from the $q^+=0$ frame, we calculated the branching ratios for $B_s\to (K,\eta,\eta')(\nu_{\ell}\bar{\nu_{\ell}},\ell^+\ell^-)$ and the LPAs for $B_s\to (K,\eta,\eta')\ell^+\ell^-$ including both short- and long-distance contributions from the QCD Wilson coefficients. Our numerical results for the nonresonant branching ratios of $B_s\to \eta^{(\prime)}(\sum\nu_{\ell}\bar{\nu_{\ell}},\mu^+\mu^-,\tau^+\tau^-)$ decays are ${\cal O}(10^{-6},10^{-7}, 10^{-8})$ in orders of magnitude, respectively. The branching ratios for the $B_s\to K(\nu_{\ell}\bar{\nu_{\ell}},\ell^+\ell^-)$ decays are at least an order of magnitude smaller than those for the $B_s\to \eta^{(\prime)}(\nu_{\ell}\bar{\nu_{\ell}},\ell^+\ell^-)$ decays. The averaged values $\la P^{K,\eta^{(\prime)}}_L\ra_\ell$ of the LPAs for $B_s\to (K,\eta^{(\prime)})\ell^+\ell^-$ without the LD contributions are $\la P^K_L\ra_\mu=\la P^\eta_L\ra_\mu=\la P^{\eta'}_L\ra_\mu=-0.98$, $\la P^K_L\ra_\tau=-0.24$, $\la P^\eta_L\ra_\tau=-0.20$ and $\la P^{\eta'}_L\ra_\tau=-0.14$, respectively. These polarization asymmetries provide valuable information on the flavor changing loop effects in the SM. Of particular interest, we estimated that the ratio ${\rm BR}(B_s\to K\mu^+\mu^-)/{\rm BR}(B_s\to \eta\mu^+\mu^-)$ differs from the ${\rm SU}_f(3)$ symmetry limit (apart from the mixing angle) by about $30\%$. Such a kind of relation may help in determining the $\eta-\eta'$ mixing angle. \ack This work was supported by the Korea Research Foundation Grant funded by the Korean Government(KRF-2008-521-C00077). \section*{References}
\section{Introduction} It is of general interest to discover a second Earth-Moon system. Detections of such an extrasolar planet with a satellite (or hypothetical binary planet systems that do not exist in the Solar System) and probing the nature of such objects will bring important information to planet (and satellite) formation theory (e.g., Williams et al. 1997, Jewitt and Sheppard 2005, Canup and Ward 2006, Jewitt and Haghighipour 2007). If a giant planet with a (perhaps Earth-size) rocky satellite were located at a certain distance from their host star, the satellite may be habitable and show vegetation, though these issues are out of the scope of this paper. It is not clear whether the IAU definition for planets in the Solar System can be applied to extrasolar planets as it is. The IAU definition in 2006 is as follows: A planet is a celestial body that (a) is in orbit around the Sun, (b) has sufficient mass so that it assumes a hydrostatic equilibrium (nearly round) shape, and (c) has cleared the neighborhood around its orbit. Regarding (c), the Earth can be called a planet, mostly because the common center of mass (COM) of the Earth-moon system is below the surface of the Earth. On the other hand, the COM of the Pluto-Charon system is located above the surfaces of these objects. Therefore, it is interesting to determine the COM position of a planet-companion system. In order to determine it, we have to know the true (not apparent) distance between the two objects. For this reason, Sato and Asada (2009) considered extrasolar mutual transits, as a complementary method of measuring not only the radii of two transiting objects but also their separation (See Sato and Asada 2009 also on detection probabilities of extrasolar mutual transits and a possible limit by Kepler Mission). As a particular case, a short separation binary, which has a rapidly orbiting companion, gives us a unique opportunity to measure the true separation of the binary, whereas a long separation one gives us only the apparent separation. Their work is very limited, however, in the sense that they assume circular orbits and also coplanar orbits as $I = 90$ degrees for both planet and moon. Clearly it is important to take account of the orbital inclination and eccentricity. The main purpose of this paper is to study effects of the orbital inclination and eccentricity of a companion on extrasolar mutual transits. Since the first detection of a transiting extrasolar planet (Charbonneau et al. 2000), photometric techniques have been successful (e.g., Deming et al. 2005 for probing atmosphere, Ohta et al. 2005, Winn et al. 2005, Gaudi \& Winn 2007, Narita et al. 2007, 2008 for measuring spin-orbit alignment angle). In addition to {\it COROT}\footnote{http://www.esa.int/SPECIALS/COROT/}, {\it Kepler}\footnote{http://kepler.nasa.gov/} is monitoring about $10^5$ stars with expected 20 ppm ($=2 \times 10^{-5}$) photometric differential sensitivity for stars of V=12. This will marginally enable the detection of a moon-size object. In fact, {\it COROT} detected a transiting super-Earth (Leger et al. 2009, Queloz et al. 2009). Sartoretti and Schneider (1999) first suggested a photometric detection of extrasolar satellites. Cabrera and Schneider (2007) developed a method based on the imaging of a planet-companion as an unresolved system (but resolved from its host star) by using planet-companion mutual transits and mutual shadows. As an alternative method, timing offsets for a single eclipse have been investigated for eclipsing binary stars as a perturbation of transiting planets around the center of mass in the presence of the third body (Deeg et al. 1998, 2000, Doyle et al. 2000). It has been recently extended toward detecting `exomoons' (Szab{\'o} et al. 2006, Simon et al. 2007, Kipping 2009a, 2009b). Sato and Asada (2009) investigated effects of mutual transits by an extrasolar planet with a companion on light curves. In particular, they studied how the effects depend on the companion's orbital velocity. Furthermore, extrasolar mutual transits were discussed as a complementary method of measuring the system's parameters such as a planet-companion's separation and thereby of identifying them as a true binary, planet-satellite system or others. Their method has analogies in classical ones for eclipsing binaries (e.g., Binnendijk 1960, Aitken 1964). A major difference is that occultation of one faint object by the other transiting a parent star causes an apparent {\it increase} in light curves, whereas eclipsing binaries make a decrease. What is more important is that, in both cases where one faint object transits the other and vice versa, changes are made in the light curves due to mutual transits even if no light emissions come from the faint objects. In a single transit, on the other hand, thermal emissions from a transiting object at lower temperature make a difference in light curves during the secondary eclipse, when the object moves behind a parent star as observed for instance for HD 209458b (Deming et al. 2005). This paper is organized as follows. In section 2, we consider effects of the orbital inclination and eccentricity of a companion on light curves. For simplicity, henceforth, such a companion orbiting around a host planet is called a `moon' even if it is not a satellite but a component of a hypothetical binary planet. In section 3, we present a formulation for parameter determinations. Some numerical examples are also presented. Section 4 is devoted to the conclusion. \section{Effects of the Orbital Eccentricity of a Transiting `Moon' on Light Curves} \subsection{Approximations and notation} The time duration of a transit, say a few hours, is much longer than the orbital period of an extrasolar planet, say a few days or greater. In one transit, the effect of the motion of the moon is much larger than that of the planet orbiting around a star. During the transit, therefore, we employ a constant velocity approximation only for the orbital motion of a planet-moon system around their host star. For a short separation case, on the other hand, we take account of the eccentric orbit of the moon, because the orbital period of such a moon around the planet may be comparable to (or shorter than) the timescale of the transit. The co-planar assumption that the orbital plane of a moon around its primary object is the same as that of the planet in orbit around the host star seems reasonable because it seems that planets are born from fragmentations of a single proto-stellar disk and thus their spins and orbital angular momentum are nearly parallel to the spin axis of the disk. Irregular satellites such as Triton, however, have significant inclinations presumably through capture processes. This requires that we should include the effect of orbital inclinations. We assume only the moon's orbital inclination, because it makes substantial effects on mutual transit light curves. Inclinations of the planet's orbital plane have been well understood and already observed (Charbonneau et al. 2000). Here we list our assumptions for clarity. \begin{itemize} \item The inclination angle of the COM of the planet and moon is fixed at 90 degrees. \item We take account of the inclination angle of the moon's orbit. \item We assume that the planet-moon COM has a constant velocity (during a transit by the planet in front of the host star). \item The longitude of ascending node of the moon equals zero. \item The planet-moon COM orbit has zero eccentricity. \item We assume no limb darkening effects. \end{itemize} For an eccentric orbit, we generally have the Kepler equation as \begin{equation} t=t_0 + \frac{T}{2\pi} (u-e \sin{u}) , \label{kepler} \end{equation} where $t_0$, $T$, $e$ and $u$ denote the time of periastron passage, orbital period, eccentricity and eccentric anomaly, respectively (e.g., Danby 1988, Roy 1988, Murray and Dermott 1999). In the following, we use the true anomaly $f$ instead of the eccentric anomaly $u$ (e.g., Danby 1988, Roy 1988, Murray and Dermott 1999). They are related by \begin{equation} \tan\frac{f}{2} = \sqrt{\frac{1+e}{1-e}} \tan\frac{u}{2} . \end{equation} The distance between the orbiting body and a focus of the ellipse is written as \begin{equation} r = \frac{a(1-e^2)}{1+e\cos f} . \end{equation} We use these equations for describing a moon orbiting around a planet. We denote the mean motion of the moon in orbit around the primary as $n_m \equiv 2\pi/T_m$, where the subscript $m$ means the moon's quantity. The subscript $p$ denotes the planet's quantity. For investigating transits, we need the transverse position $x$ and velocity $v$ of each object. We denote those of the COM for planet-moon systems as $x_{CM}$ and $v_{CM}$, respectively, where the origin of $x$ is chosen as the center of the star. We assume $v_{CM}$ as constant during the transit. The position and velocity of each planet with mass $M_p$ and $M_m$ in the planet-moon system as $x_i$ and $v_i$ ($i=p, m$), respectively. The direction of the observer's line of sight is specified by the argument of pericenter as an angle denoted by $\omega_m$ (See also Fig. $\ref{config}$). We express the transverse position as \begin{eqnarray} x_{CM} &=& v_{CM} (t-t_{CM}) , \\ \label{xCM} x_p &=& x_{CM}+a_p [ (\cos u - e_m) \cos\omega_m + \sqrt{1-e_m^2} \sin u \sin\omega_m ] , \label{x1-u} \\ x_m &=& x_{CM}-a_m [ (\cos u - e_m) \cos\omega_m + \sqrt{1-e_m^2} \sin u \sin\omega_m ] , \label{x2-u} \end{eqnarray} where the semimajor axis of the orbit of each object around their COM is denoted by $a_i$, and $t_{CM}$ means the time when the binary's common center of mass passes in front of the center of the host star. In terms of the true anomaly, they are rewritten as \begin{eqnarray} x_p &=& x_{CM}+a_p \frac{(1-e_m^2) \cos(f+\omega_m)}{1 + e_m \cos f} , \label{x1-f} \\ x_m &=& x_{CM}-a_m \frac{(1-e_m^2) \cos(f+\omega_m)}{1 + e_m \cos f} . \label{x2-f} \end{eqnarray} (See Table 1 for a list of parameters and their definition). The azimuthal velocity of the secondary object around the primary is \begin{eqnarray} V_f &=& r \frac{df}{dt} \nonumber\\ &=& a_{pm}n_m \frac{1+e_m\cos f}{\sqrt{1-e_m^2}} , \label{Vf} \end{eqnarray} where $f$ denotes the true anomaly (Murray and Dermott 2000). Here, $a_{pm}$ and $e_m$ denote the semimajor axis and eccentricity of the moon's orbit with respect to the host planet. We assume that both the eccentricity of COM orbit vanishes and the inclination angle of COM equals 90 degrees. Hence we can avoid a careful treatment of ``sky-projected transverse position''. \subsection{Transits in light curves} We denote the intrinsic stellar luminosity as $L$. The apparent luminosity $L^{'}$ due to mutual transits is expressed as \begin{equation} L^{'} = L \times \frac{S-\Delta S}{S} , \label{L} \end{equation} where $S=\pi R_s^2$, $S_p=\pi R_p^2$, $S_m=\pi R_m^2$, $\Delta S=S_p+S_m-S_{pm}$. Here, $R_s$, $R_p$ and $R_m$ denote the radii of the host star, planet and moon, and $S_{pm}$ denotes the area of the apparent overlap between them, which is seen from the observer. Without loss of generality, we assume that the primary is larger than the secondary as $R_p \geq R_m$. \subsection{Effects on light curves} We investigate light curves by mutual transits due to planet-moon systems. The orbital velocity is of the order of $a_{pm}n_m$. Therefore, we have two cases; $v_{CM} < a_{pm}n_m$ and $v_{CM} > a_{pm}n_m$. The dimensionless ratio of the moon's orbital velocity to the planet's one is defined as \begin{equation} W \equiv \frac{a_{pm}n_m}{v_{CM}} . \label{W} \end{equation} If $v_{CM} < a_{pm} n_m$, we call it a fast case. If $v_{CM} > a_{pm} n_m$, we call it a slow one. The Earth-Moon ($W=0.03$), Jupiter-Ganymede ($W=0.8$) and Jupiter-Io ($W=1.3$) systems represent slow, marginal and fast cases, respectively. Figure $\ref{schematic}$ shows a schematic light curve by mutual transits. Fig. $\ref{lightcurve-1}$ shows a slow case in circular motion, where we assume $R_s: R_p: R_m = 20: 2: 1$. We assume also the same mass density for the two transiting objects and hence obtain $a_p: a_m = 1: 8$. Eccentric orbit cases ($W=6$ and $e_m=0.3$) are shown by Figs. $\ref{lightcurve-2}$, $\ref{lightcurve-3}$ and $\ref{lightcurve-4}$ ($\omega_m=\pi/2$, $0$ and $\pi/4$, respectively). Some parameters are chosen so that effects in the figures can be distinguished by eye, though such an event is unlikely to be detected by current observations as discussed later. For generating the ingress and egress of the various parts of the lightcurve, we do not use a linear interpolation but compute numerically the apparent overlap area between the objects. Here we assume the same configuration except for the observer's line of sight. For simplicity, we take $t_0 = t_{CM} = 0$ in these figures. These figures show also the transverse positions of transiting objects with time, which would help us to understand the chronological changes in the light curves. In particular, it can be understood that such characteristic patterns appear only when two objects are in front of the star and one of them transits (or occults) the other. \section{Formulation for Parameter Determinations by Transit Method} \subsection{Parameter determinations from transit observations alone} In all the above cases, the amount of decrease in light curves or the magnitude of fluctuations gives the ratios among the radii of the star and two faint objects $(R_s, R_p, R_m)$. The decrease ratios in the apparent brightness due to transits by the planet and moon are written as \begin{eqnarray} \Delta_p &=& \left(\frac{R_p}{R_s}\right)^2 , \label{Delta1} \\ \Delta_m &=& \left(\frac{R_m}{R_s}\right)^2 . \label{Delta2} \end{eqnarray} The stellar radius $R_s$ (and mass $M_S$) are known for instance by its spectral type. Hence, the radii are expressed in terms of observables $R_s$, $\Delta_p$ and $\Delta_m$ as \begin{eqnarray} R_p &=& R_s \sqrt{\Delta_p} , \label{r1} \\ R_m &=& R_s \sqrt{\Delta_m} . \label{r2} \end{eqnarray} We define the ratio between the brightness changes by the two objects as \begin{equation} \Delta \equiv \frac{\Delta_m}{\Delta_p} . \end{equation} \noindent {\bf Circular Orbit and Orbital Inclination: }\\ First, we discuss a circular orbit in order to simply explain our idea. For more rigorous treatment of eccentric orbits, please see below, where we will finally give expressions for determining the separation $a_{pm}$. Behaviors of apparent light curves depend on $W$. Therefore, $a_{pm}n_m$ (as its ratio to $v_{CM}$) can be obtained (Sato and Asada 2009). Seager and Mall{\'e}n-Ornelas (2003) presents an analytic solution of parameter determinations for a single transit in the circular orbit case. Their solution can be used for our case of mutual transits by a planet and moon in front of a host star. In our case, their equations are rewritten as follows. The duration of the `flat part' of the transit ($t_{Fm}$) is described by \begin{equation} \sin\left( \frac{t_{Fm} \pi}{T_m} \right) = \frac{1}{a_{pm} \sin I_m} \sqrt{(R_p-R_m)^2 - (a_{pm} \cos I_m)^2} , \end{equation} whereas the total transit duration ($t_{Tm}$) is done by \begin{equation} \sin\left( \frac{t_{Tm} \pi}{T_m} \right) = \frac{1}{a_{pm} \sin I_m} \sqrt{(R_p+R_m)^2 - (a_{pm} \cos I_m)^2} , \label{sintT} \end{equation} where $I_m$ denotes the orbital inclination angle of the moon. By combining these equations, the impact parameter ($b_{pm}$) can be derived as \begin{eqnarray} b_{pm} &\equiv& \frac{a_{pm}}{R_p} \cos I_m \nonumber\\ &=& \sqrt{\frac{(1-\sqrt{\Delta})^2 - [\sin^2(t_{Fm}\pi/T_m)/\sin^2(t_{Tm}\pi/T_m)] (1+\sqrt{\Delta})^2}{1-[\sin^2(t_{Fm}\pi/T_m)/\sin^2(t_{Tm}\pi/T_m)]}} . \end{eqnarray} The ratio $a_{pm}/R_p$ can be derived directly from Eq. ($\ref{sintT}$) as \begin{equation} \frac{a_{pm}}{R_p} = \sqrt{\frac{(1+\sqrt{\Delta})^2 - b_{pm}^2 [1-\sin^2(t_{Tm}\pi/T_m)]} {\sin^2(t_{Tm}\pi/T_m)}} . \end{equation} Therefore, one can obtain $a_{pm}$ as $R_s \times (R_p/R_s) \times (a_{pm}/R_p)$. With $a_{pm}$ and $n_m$ in hand, one can thus estimate the total mass of the binary by $GM_{tot}=n_m^2 a_{pm}^3$ from Kepler's third law, where $G$ denotes the gravitational constant. The orbital velocity $a_{pm}n_m$ gives the mutual force between the binary. If we assume also that the mass density is common for two objects constituting the binary (this may be reasonable especially for similar size objects as $R_p \sim R_m$), each mass is determined as $M_p = R_p^3 (R_p^3 + R_m^3)^{-1} M_{tot}$ and $M_m = R_m^3 (R_p^3 + R_m^3)^{-1} M_{tot}$, respectively. Therefore, the orbital radius of each body around the COM is obtained as $a_p = R_m^3 (R_p^3 + R_m^3)^{-1} a_{pm}$ and $a_m = R_p^3 (R_p^3 + R_m^3)^{-1} a_{pm}$, respectively. At this point, importantly, the two objects can be identified as a true binary ($a_p>R_p$) or planet-satellite system ($a_p<R_p$). However, a gaseous giant planet with a rocky satellite would exhibit largely different densities and this may be one of the most likely scenarios. In a slow spin case, on the other hand, the apparent separation $a_{\perp}$ (normal to our line of sight) is determined as $a_{\perp} = T_{12} v_{CM}$ from measuring the time lag $T_{12}$ between the first and second transits because $v_{CM}$ is known above (Sato and Asada 2009). \noindent {\bf Eccentric Orbit and Edge-on Case:}\\ Henceforth, we take account of the orbital eccentricity of a moon for an edge-on case. In this case, intervals between neighboring ``hills'' are not constant because of the eccentricity. However, a time duration between three successive ``hills'' is nothing but the orbital period of the moon. Therefore, one can measure the period $T_m$. We obtain $n_m$ as \begin{equation} n_m=\frac{2\pi}{T_m} . \label{omega} \end{equation} A key idea for determining the eccentricity is as follows. As for timescales, we have two observable ratios as $T_2/T_1$ and $T_{12}/T_{21}$. The former is the ratio between the widths of neighboring hills, whereas the latter is that between the transit intervals. On the other hand, we have two additional parameters $e_m$ and $\omega_m$ to be determined. Importantly, the number of measurable ratios is the same as that of the parameters that we wish to determine. In principle, therefore, the above two ratios may allow us to determine the two parameters $e_m$ and $\omega_m$, separately. This will be discussed in detail below. To be more precise, the full width of a ``hill'' at top and bottom are expressed as (See also Figure $\ref{schematic}$) \begin{eqnarray} T_{top} &=& \frac{2(R_p-R_m)}{V_f}, \label{Ttop} \\ T_{bottom} &=& \frac{2(R_p+R_m)}{V_f} . \label{Tbottom} \end{eqnarray} Only for symmetric binaries ($R_p=R_m$), we have $T_{top}=0$ and thus true spikes. Otherwise, truncated spikes (or ``hills'') appear. For the primary transit, where the moon moves in front of the planet, we have $f=\pi/2 - \omega_m$. {}From Eqs. ($\ref{Vf}$), ($\ref{Ttop}$) and ($\ref{Tbottom}$), therefore, we obtain \begin{eqnarray} T_{1 top} &=& \frac{2(R_p-R_m)}{a_{pm}n_m} \frac{\sqrt{1-e_m^2}}{1+e_m\sin\omega_m} , \label{T1top} \\ T_{1 bottom} &=& \frac{2(R_p+R_m)}{a_{pm}n_m} \frac{\sqrt{1-e_m^2}}{1+e_m\sin\omega_m} . \label{T1bottom} \end{eqnarray} For the circular orbit $e_m=0$, the second factors in the R.H.S. of these expressions become the unity and the first factors recover the case for $e_m=0$ (See Sato and Asada 2009). For the secondary transit, where the moon moves behind the planet, we have $f=3\pi/2 - \omega_m$. From Eqs. ($\ref{Vf}$), ($\ref{Ttop}$) and ($\ref{Tbottom}$), therefore, we obtain \begin{eqnarray} T_{2 top} &=& \frac{2(R_p-R_m)}{a_{pm}n_m} \frac{\sqrt{1-e_m^2}}{1-e_m\sin\omega_m} , \label{T2top} \\ T_{2 bottom} &=& \frac{2(R_p+R_m)}{a_{pm}n_m} \frac{\sqrt{1-e_m^2}}{1-e_m\sin\omega_m} . \label{T2bottom} \end{eqnarray} We immediately obtain from Eqs. ($\ref{T1top}$), ($\ref{T1bottom}$), ($\ref{T2top}$) and ($\ref{T2bottom}$), \begin{eqnarray} T_r &\equiv& \frac{T_{2 top}}{T_{1 top}} \nonumber\\ &=& \frac{T_{2 bottom}}{T_{1 bottom}} \nonumber\\ &=&\frac{1+e_m\sin\omega_m}{1-e_m\sin\omega_m} . \label{Tr} \end{eqnarray} Equation ($\ref{Tr}$) is rewritten as \begin{eqnarray} e_m\sin\omega_m &=& \frac{T_r-1}{T_r+1} \nonumber\\ &\equiv& T_R, \label{ecosPsi} \end{eqnarray} where the R.H.S. can be determined by observations alone. Once either $e_m$ or $\omega_m$ is known, Eq. ($\ref{ecosPsi}$) determines the other. Next, we consider the time interval between the primary and secondary transits. In order to compute such an interval, one can use the Kepler's second law (the constant areal velocity). After lengthy but straightforward calculations, the area swept from the primary transit ($f=\pi/2 - \omega_m$) till the secondary ($f=3\pi/2 - \omega_m$) becomes \begin{eqnarray} S_{12}&=& a_mb_m \times \left[ \frac{\pi}{2}+\arcsin H +\frac12 \sin(2\arcsin H) \right] , \label{S12} \end{eqnarray} where $b_m$ denotes the semiminor axis of the moon's elliptic orbit and $H$ is defined as \begin{equation} H=e_m\sqrt{\frac{\cot^2\omega_m}{1-e_m^2+\cot^2\omega_m}} . \label{H} \end{equation} By using Eq. ($\ref{ecosPsi}$) in Eq. ($\ref{H}$) for eliminating $\cot\omega_m$, we obtain \begin{equation} H=\sqrt{\frac{e_m^2-T_R^2}{1-T_R^2}} . \label{H2} \end{equation} It is convenient to use $T_{12}/T$ instead of $T_{12}/T_{21}$, because $T_{12}/T$ gives a simpler expression than $T_{12}/T_{21}$, which is used in the above explanation of the key idea. The total area is $S=\pi a_m b_m$. Therefore, we find \begin{eqnarray} \frac{T_{12}}{T}&=&\frac{S_{12}}{S} \nonumber\\ &=&\frac12 +\frac{1}{\pi}\arcsin H +\frac{1}{\pi} H \sqrt{\frac{1-e_m^2}{1-T_R^2}} . \label{Tratio} \end{eqnarray} This includes $e_m$ without $\omega_m$. Hence by using this relation, one can determine separately the eccentricity by measuring the time intervals. We should note that Eq. ($\ref{Tratio}$) is valid even for a general case with a certain inclination angle. Here, we consider a small eccentricity approximation, which may be useful for a quicker estimation of the parameters. For small $e_m$, Eq. ($\ref{H}$) is expanded as \begin{equation} H=e_m \cos\omega_m+O(e_m^3) . \label{H-expansion} \end{equation} Substitution of this into Eq. ($\ref{Tratio}$) gives us \begin{eqnarray} \frac{T_{12}}{T_m} &=&\frac12 +\frac{2e_m}{\pi}\cos\omega_m +O(e_m^3) . \label{Tratio-expansion} \end{eqnarray} The correction to the circular case ($T_{12}/T_m=1/2$) is $2e_m \cos\omega_m/\pi \sim 0.7 e_m \cos\omega_m$. Even for $\omega_m\sim 0$ for instance, it leads to seven percents for $e_m=0.1$. We define the difference between $T_{12}$ and $T_{21}$ as \begin{equation} \delta T \equiv T_{12}-T_{21} . \label{deltaT} \end{equation} Replacement $\omega_m$ by $\omega_m+\pi$ changes Eq. ($\ref{Tratio-expansion}$) into \begin{eqnarray} \frac{T_{21}}{T_m} &=&\frac12 -\frac{2e_m}{\pi}\cos\omega_m +O(e_m^3) . \label{Tratio-expansion2} \end{eqnarray} By using Eqs. ($\ref{Tratio-expansion}$) and ($\ref{Tratio-expansion2}$), we obtain \begin{equation} e_m \cos\omega_m = \frac{\pi \delta T}{4 T_m} + O(e^3) . \label{deltaT2} \end{equation} Let us consider every cases separately. (I) If and only if the L.H.S. of Eqs. ($\ref{ecosPsi}$) and ($\ref{deltaT2}$) vanish, $e_m$ does. This means a circular motion and thus the observer's direction $\omega_m$ becomes meaningless. (II) For a case when the L.H.S. of Eq. ($\ref{ecosPsi}$) vanishes but that of Eq. ($\ref{deltaT2}$) does not, we find $e_m \neq 0$ and $\sin\omega_m = 0$, namely \begin{equation} \omega_m=0 \: (\mbox{mod}\: \pi). \end{equation} Eq. ($\ref{deltaT2}$) immediately gives \begin{equation} e_m=\frac{\pi \delta T}{4 T_m} . \end{equation} Hence, the orbital eccentricity is determined. (III) If the L.H.S. in Eq. ($\ref{deltaT2}$) vanishes but that in ($\ref{ecosPsi}$) does not, we find $e_m \neq 0$ and $\cos\omega_m = 0$, namely \begin{equation} \omega_m=\frac{\pi}{2} \: (\mbox{mod}\: \pi). \end{equation} Eq. ($\ref{ecosPsi}$) immediately gives \begin{equation} e_m=T_R . \end{equation} Hence, the orbital eccentricity is measured. (VI) A general case in which the L.H.S. of neither Eqs. ($\ref{ecosPsi}$) nor ($\ref{deltaT2}$) vanish: By dividing Eq. ($\ref{deltaT2}$) by Eq. ($\ref{ecosPsi}$), we obtain \begin{equation} \cot\omega_m = \frac{\pi}{4} \frac{\delta T}{T_m} \frac{1}{T_R} , \label{tanPsi} \end{equation} where the R.H.S. can be determined by observations alone and hence this equation gives us the observer's direction $\omega_m$. By substituting the determined $\omega_m$ into Eq. ($\ref{ecosPsi}$), one can find the value of the eccentricity. Up to this point, $e_m$ and $\omega_m$ both are determined. Eqs. $(\ref{T1top})$ and $(\ref{T1bottom})$ are rewritten as \begin{eqnarray} a_{pm} &=& \frac{T_m (R_p-R_m)}{\pi T_{1top}} \frac{\sqrt{1-e_m^2}}{1+e_m\sin\omega_m} \nonumber\\ &=& \frac{T_m (R_p-R_m) (T_{1top}+T_{2top}) \sqrt{1-e_m^2}} {2\pi T_{1top}T_{2top}} \nonumber\\ &=& \frac{(R_p-R_m)}{2\pi} \left(\frac{T_m}{T_{1top}}+\frac{T_m}{T_{2top}}\right) \sqrt{1-e_m^2} , \label{a1top} \\ a_{pm} &=& \frac{T_m (R_p+R_m)}{\pi T_{1bottom}} \frac{\sqrt{1-e_m^2}}{1+e_m\sin\omega_m} \nonumber\\ &=& \frac{T_m (R_p+R_m) (T_{1bottom}+T_{2bottom}) \sqrt{1-e_m^2}} {2\pi T_{1bottom}T_{2bottom}} \nonumber\\ &=& \frac{(R_p+R_m)}{2\pi} \left(\frac{T_m}{T_{1bottom}}+\frac{T_m}{T_{2bottom}}\right) \sqrt{1-e_m^2} , \label{a1bottom} \end{eqnarray} respectively, where we used Eq. ($\ref{ecosPsi}$) If and only if $R_p=R_m$, we obtain $T_{1top} = T_{2top} = 0$ and Eq. ($\ref{a1top}$) thus becomes undetermined, whereas Eq. ($\ref{a1bottom}$) is still well-defined. When one wishes to consider the secondary transit instead of the first, one can use \begin{eqnarray} a_{pm} &=& \frac{T_m (R_p-R_m)}{\pi T_{2top}} \frac{\sqrt{1-e_m^2}}{1-e_m\sin\omega_m} , \label{a2top} \\ a_{pm} &=& \frac{T_m (R_p+R_m)}{\pi T_{2bottom}} \frac{\sqrt{1-e_m^2}}{1-e_m\sin\omega_m} . \label{a2bottom} \end{eqnarray} They are obtained by replacing $\omega_m$ with $\omega_m+\pi$ in Eqs. ($\ref{a1top}$) and ($\ref{a1bottom}$). By noting $T_{1top} (1+T_R) = T_{2bottom} (1-T_R)$ and $T_{2top} (1+T_R) = T_{2bottom} (1-T_R)$, one can show that Eqs. ($\ref{a2top}$) and ($\ref{a2bottom}$) agree with Eqs. ($\ref{a1top}$) and ($\ref{a1bottom}$), respectively. By using one of these expressions, we can thus measure the semimajor axis of the eccentric orbit. In terms of the decrease in apparent brightness, Eqs. ($\ref{a1top}$) and ($\ref{a1bottom}$) are written as \begin{eqnarray} \frac{a_{pm}}{R_s} &=& \frac{(\sqrt{\Delta_p}-\sqrt{\Delta_m})}{2\pi} \left(\frac{T_m}{T_{1top}}+\frac{T_m}{T_{2top}}\right) \sqrt{1-e_m^2} , \label{a1top2} \\ \frac{a_{pm}}{R_s} &=& \frac{(\sqrt{\Delta_p}+\sqrt{\Delta_m})}{2\pi} \left(\frac{T_m}{T_{1bottom}}+\frac{T_m}{T_{2bottom}}\right) \sqrt{1-e_m^2} , \label{a1bottom2} \end{eqnarray} where we used Eqs. ($\ref{r1}$) and ($\ref{r2}$). Determination of the semimajor axis is sensitive to measurement errors in the widths of the hills. This statement can be proven by using Eq. ($\ref{a1bottom2}$). For simplicity, we assume $\Delta_p \sim \Delta_m \sim \Delta$ and $T_{1bottom} \sim T_{2bottom} \sim T_{bottom}$, so that Eq. ($\ref{a1bottom2}$) can be reduced to \begin{equation} a_{pm} \sim \frac{2}{\pi} R_s \sqrt{\Delta} \frac{T_m}{T_{bottom}} \sqrt{1-e_m^2} . \label{asim} \end{equation} The logarithmic derivative of this becomes \begin{equation} \frac{da_{pm}}{a_{pm}} \sim \frac{dR_s}{R_s} + \frac{d\Delta}{2\Delta} + \frac{dT_m}{T_m} - \frac{dT_{bottom}}{T_{bottom}} - \frac{e_mde_m}{(1-e_m^2)} . \label{daa} \end{equation} We focus on $dT_m/T_m$ and $dT_{bottom}/T_{bottom}$, because we can expect much more accurate measurements for $R_s$ and $\Delta$, and the last term involving $e_m$ may be relatively small ($e_m<1$ and $de_m<1$). Time resolution in observations seems $dT_m \sim dT_{bottom}$, while $T_m \gg T_{bottom}$. Therefore, $dT_m/T_m \ll dT_{bottom}/T_{bottom}$, which means that accurate measurements of $T_{bottom}$ are crucial for the determination of $a_{pm}$. Figure $\ref{flow-chart}$ shows a flow chart of the parameter determinations that are discussed above. The above formulation for parameter determinations actually recovers the correct values in Figs. $\ref{lightcurve-2}$-$\ref{lightcurve-4}$. In the numerical examples, the original parameters are retrieved within twenty percents. \noindent {\bf Eccentric Orbit and Orbital Inclination:}\\ Figure $\ref{lightcurve-inclination}$ shows a difference between light curves for the edge-on ($I_m = 90$ deg.) and an inclination case ($I_m = 88$ deg.). Let $z$-axis denote the axis normal to the $x$-axis on the celestial sphere. We define the distance of a `moon' from the $z$-axis at the initial time of the mutual transit as \begin{equation} s_b = \sqrt{(R_p+R_m)^2 - \left(\frac{a_{pm}(1-e_m^2)}{1+e_m\cos f}\right)^2 \cos^2 I_m} , \label{sb} \end{equation} where the subscript $b$ means that the quantity is related with $T_{1 bottom}$ and $T_{2 bottom}$ as shown below (See also Fig. $\ref{definition}$). Similarly, when the `flat part' of the spike in light curves starts (or ends), we define the distance of a moon from the $z$-axis at this epoch as \begin{equation} s_t = \sqrt{(R_p-R_m)^2 - \left(\frac{a_{pm}(1-e_m^2)}{1+e_m\cos f}\right)^2 \cos^2 I_m} , \label{st} \end{equation} where the subscript $t$ means that the quantity is related with $T_{1 top}$ and $T_{2 top}$ as shown below. Therefore, we obtain the duration $T_{top}$ as \begin{equation} T_{top} = \frac{2s_t}{V_f} , \end{equation} where $V_f$ is given by Eq. ($\ref{Vf}$). For the primary transit ($f=\pi/2 - \omega_m$), we thus obtain \begin{equation} T_{1 top} = \frac{2s_{t1} \sqrt{1-e_m^2}}{a_{pm}n_m (1+e_m\sin\omega_m)} , \label{T1top-I} \end{equation} whereas for the secondary ($f=3\pi/2 - \omega_m$), we have \begin{equation} T_{2 top} = \frac{2s_{t2} \sqrt{1-e_m^2}}{a_{pm}n_m (1-e_m\sin\omega_m)} . \label{T2top-I} \end{equation} Here, we define $s_{t1}$ and $s_{t2}$ as \begin{eqnarray} s_{t1}&=& \sqrt{(R_p-R_m)^2 - \left(\frac{a_{pm}(1-e_m^2)}{1+e_m\sin\omega_m}\right)^2 \cos^2 I_m} , \label{st1} \\ s_{t2}&=& \sqrt{(R_p-R_m)^2 - \left(\frac{a_{pm}(1-e_m^2)}{1-e_m\sin\omega_m}\right)^2 \cos^2 I_m} . \label{st2} \end{eqnarray} In the similar manner, we obtain the width of the spikes at the bottom as \begin{equation} T_{bottom} = \frac{2s_b}{V_f} . \end{equation} For the primary transit ($f=\pi/2 - \omega_m$), we thus obtain \begin{equation} T_{1 bottom} = \frac{2s_{b1} \sqrt{1-e_m^2}}{a_{pm}n_m (1+e_m\sin\omega_m)} , \label{T1bottom-I} \end{equation} whereas for the secondary ($f=3\pi/2 - \omega_m$), we have \begin{equation} T_{2 bottom} = \frac{2s_{b2} \sqrt{1-e_m^2}}{a_{pm}n_m (1-e_m\sin\omega_m)} , \label{T2bottom-I} \end{equation} where we define $s_{b1}$ and $s_{b2}$ as \begin{eqnarray} s_{b1}&=& \sqrt{(R_p+R_m)^2 - \left(\frac{a_{pm}(1-e_m^2)}{1+e_m\sin\omega_m}\right)^2 \cos^2 I_m} , \label{sb1} \\ s_{b2}&=& \sqrt{(R_p+R_m)^2 - \left(\frac{a_{pm}(1-e_m^2)}{1-e_m\sin\omega_m}\right)^2 \cos^2 I_m} . \label{sb2} \end{eqnarray} Because of the orbital inclination, we have to consider $T_{top}$ and $T_{bottom}$, separately. We define the ratios as \begin{equation} T_{r top} \equiv \frac{T_{2 top}}{T_{1 top}} , \end{equation} and \begin{equation} T_{r bottom} \equiv \frac{T_{2 bottom}}{T_{1 bottom}} . \end{equation} Substitutions of Eqs. ($\ref{T1top-I}$), ($\ref{T2top-I}$), ($\ref{T1bottom-I}$), ($\ref{T2bottom-I}$) into these ratios lead to \begin{eqnarray} T_{r top}&=& \frac{s_{t2}}{s_{t1}} \frac{1+e_m\sin\omega_m}{1-e_m\sin\omega_m} , \label{Trtop-I} \\ T_{r bottom}&=& \frac{s_{b2}}{s_{b1}} \frac{1+e_m\sin\omega_m}{1-e_m\sin\omega_m} . \label{Trbottom-I} \end{eqnarray} For the edge-on case ($I_m = 90$ deg.), we obtain $s_{t2}/s_{t1} = s_{b2}/s_{b1} = 1$. Then, we have $T_{r top} = T_{r bottom}$. For a general case ($I_m \neq 90$ deg.), on the other hand, we find $T_{r top} \neq T_{r bottom}$. We thus expect that a ratio between them will give us the information about the orbital inclination. The ratio is \begin{equation} \frac{T_{r top}}{T_{r bottom}} = \frac{s_{t2}}{s_{t1}} \frac{s_{b1}}{s_{b2}} . \label{Tratio-I} \end{equation} The L.H.S. can be measured by observations. There are four unknown quantities $a_{pm}$, $e_m$, $I_m$ and $\omega_m$. We have four equations of ($\ref{Trtop-I}$), ($\ref{Trbottom-I}$), ($\ref{Tratio-I}$) and the last one that can be chosen out of ($\ref{T1top-I}$), ($\ref{T2top-I}$), ($\ref{T1bottom-I}$) and ($\ref{T2bottom-I}$). Therefore, one can determine the quantities $a_{pm}$, $e_m$, $I_m$ and $\omega_m$ by using these equations for observations. For practical observations, data fittings at the slope of light curves are used, instead of transit durations, for determinations of the orbital inclination angle (Charbonneau et al. 2000). Nevertheless, an analytic solution is necessarily worthwhile to understand the properties of a given physical system, even when numerical fits are in practice the best way to determine the system parameters. A partial transit occurs if the apparent impact parameter of the moon is in $(R_p-R_m, R_p+R_m)$. For the primary transit ($f=\pi/2 - \omega_m$), it occurs if the orbital inclination angle satisfies \begin{equation} \frac{R_p-R_m}{a_{pm}} \frac{1+e_m\sin\omega_m}{1-e_m^2} < \cos I_m < \frac{R_p+R_m}{a_{pm}} \frac{1+e_m\sin\omega_m}{1-e_m^2} . \end{equation} For the secondary one ($f=3\pi/2 - \omega_m$), the condition of a partial transit becomes \begin{equation} \frac{R_p-R_m}{a_{pm}} \frac{1-e_m\sin\omega_m}{1-e_m^2} < \cos I_m < \frac{R_p+R_m}{a_{pm}} \frac{1-e_m\sin\omega_m}{1-e_m^2} . \end{equation} Such a partial transit by a moon orbiting a host planet produces a `U'-shaped spike in light curves (See Fig. $\ref{lightcurve-partial}$). \subsection{Timescales and brightness changes} We have presented a formalism for parameter determinations. Before closing this section, let us make brief comments on typical timescales and amplitudes in the brightness changes. The timescale of a brightness change due to a giant planet is about \begin{equation} \frac{R_p}{a_{pm}n_m} \sim 5 \times 10^3 \left( \frac{R_p}{5 \times 10^4 \mbox{km}}\frac{10 \mbox{km/s}}{a_{pm}n_m} \right) \mbox{sec.} \end{equation} Therefore, {\it detections} of such fluctuations due to mutual transits of extrasolar planet-moon systems require frequent observations, say every hour. Furthermore, higher frequency (e.g., every ten minutes) is necessary for parameter {\it estimations} of the system. Let us mention a connection of the present result with space telescopes in operation. Decrease in apparent luminosity due to the secondary planet is $O(R_m^2/R_s^2)$. Besides the time resolution (or observation frequency) and mission lifetimes, detection limits by {\it COROT} with the achieved accuracy of photometric measurements (700 ppm in one hour) could put $R_m/R_s \sim 2 \times 10^{-2}$. The nominal integration time is 32 sec. but co-added over 8.5 min. except for 1000 selected targets for which the nominal sampling is preserved. By the {\it Kepler} mission with expected 20 ppm differential sensitivity for solar-like stars with $m_V=12$, the lower limit will be reduced to $R_m/R_s \sim 4 \times 10^{-3}$. An analogy of the Earth-Moon ($R_m/R_s \sim 2.5 \times 10^{-3}$, $W \sim 0.03$) and Jupiter-Ganymede ($R_m/R_s \sim 4 \times 10^{-3}$, $W \sim 0.8$) will be marginally detectable. Observations both with high frequency (at least during the time of transits) and with good photometric sensitivity are desired for future detections of mutual transits. \subsection{Dynamical limit and constraints on W} For Roche limit, we have \begin{equation} a_{pm} < \beta R_H , \end{equation} where $\beta$ denotes a numerical coefficient $0 < \beta <1$ and $R_H$ is Hill radius (See Domingos et al. 2006 for more detailed stability arguments by numerical computations). For simplicity, we assume $M_s \gg M_p \gg M_m$. Then, we have $a_{pm} \approx a_m$ and $R_H$ is approximated as \begin{equation} R_H = \left( \frac{M_p}{3M_s} \right)^{1/3} d_p , \end{equation} where $M_s$ denotes a host star mass and $d_p$ denotes the orbital radius of a planet orbiting the star. Kepler's third law gives $a_{pm}n_m$ and $v_{CM}$ as \begin{eqnarray} a_{pm}n_m &=& \sqrt{\frac{GM_p}{a_m}} , \\ v_{CM} &=& \sqrt{\frac{GM_s}{d_p}} . \end{eqnarray} Combining these relations, therefore, we find \begin{equation} W > 3^{1/6} \beta^{-1/2} \left( \frac{M_p}{M_s} \right)^{1/3} . \label{W>} \end{equation} If one assumes $M_p \sim M_J$ (Jupiter mass), we obtain $W > 10^{-1}$. This lower bound is less severe. On the other hand, there is a stringent constraint that the moon's closest approach $r_{min}$ cannot be within the planetary radius. We thus have \begin{equation} a_m > \frac{r_{min}}{1-e_m} . \end{equation} This leads to \begin{equation} W < \sqrt{ \frac{M_p}{M_s} \frac{d_p}{r_{min}} (1-e_m)} . \end{equation} If one assumes the jovian mass and radius ($R_J$), this is rewritten as \begin{equation} W < 1.4 \left( \frac{d_p}{1 \mbox{AU}} \right)^{1/2} \left(\frac{M_{\odot}}{M_s}\right)^{1/2} \left(\frac{M_J}{M_p}\right)^{1/2} \left( \frac{R_J}{r_{min}} \right)^{1/2} (1-e_m)^{1/2} . \label{W<} \end{equation} For $W=6$ and $e_m=0.3$, we obtain $d_p > 25$ AU. Namely, a planet with a long orbital period $T_p > 125$ years is required. Kepler mission for several years is unlikely to see a transit by such a long period planet. Next, we consider a constraint that a mutual transit can occur. The transit duration for a planet in circular orbit is \begin{equation} D \sim \frac{T_p R_s}{\pi d_p} , \end{equation} where we assume the maximum duration by taking the vanishing impact parameter (See Seager and Mall{\'e}n-Ornelas for a more accurate form). As this limiting case for the Roche limit, we obtain the fastest case of a `moon' as \begin{equation} a_m \sim \beta \left( \frac{M_p}{3 M_s} \right)^{1/3} d_p . \end{equation} Using the Kepler's third law, this leads to \begin{equation} T_m \sim \left(\frac{\beta^3}{3}\right)^{1/2} T_p . \end{equation} For our fast case, we require $T_m < D$, which gives a bound on $\beta$ as \begin{equation} \beta < 3^{1/3} \left( \frac{R_s}{\pi d_p} \right)^{2/3} . \end{equation} We substitute this into $\beta$ of Eq. ($\ref{W>}$) so that we can obtain \begin{eqnarray} W&>&\left( \frac{\pi d_p}{R_s} \right)^{1/3} \left( \frac{M_p}{M_s} \right)^{1/3} \nonumber\\ &\sim& 1.5 \left( \frac{d_p}{d_J} \right)^{1/3} \left( \frac{R_{\odot}}{R_s} \right)^{1/3} \left( \frac{M_p}{M_J} \right)^{1/3} \left( \frac{M_{\odot}}{M_s} \right)^{1/3} \nonumber\\ &\sim& 0.9 \left( \frac{T_p}{1 \mbox{year}} \right)^{2/9} \left( \frac{R_{\odot}}{R_s} \right)^{1/3} \left( \frac{M_p}{M_J} \right)^{1/3} \left( \frac{M_{\odot}}{M_s} \right)^{1/3} , \label{W>2} \end{eqnarray} where $d_J$ denotes the mean orbital radius of the Jupiter. On the other hand, the dynamical arguments put an upper bound by Eq. ($\ref{W<}$). This is rewritten as \begin{equation} W < 1.4 \left( \frac{T_p}{1 \mbox{year}} \right)^{1/3} \left(\frac{M_{\odot}}{M_s}\right)^{1/2} \left(\frac{M_J}{M_p}\right)^{1/2} \left( \frac{R_J}{r_{min}} \right)^{1/2} (1-e_m)^{1/2} . \label{W<2} \end{equation} Therefore, there exists a narrow band as shown by Fig. $\ref{band}$. Outside this range, the proposed method cannot work. \section{Conclusion} We have shown that light curves by mutual transits of an extrasolar planet with a `moon' depend on the moon's orbital eccentricity, especially for small separation (fast) cases, in which occultation of one faint object by the other transiting a parent star causes an apparent increase in light curves and such characteristic fluctuations with the same height repeatedly appear. We have also presented a formulation for determining the parameters such as the orbital eccentricity, inclination, semimajor axis and the direction of the observer's line of sight. This will be useful for probing the nature of the transiting planet-moon system. When actual light curves are analyzed, we should incorporate (1) photometric corrections such as limb darkenings, and (2) perturbations as three (or more)-body gravitating interactions (e.g., Danby 1988, Murray and Dermott 2000). We would like to thank the referee for invaluable comments on the manuscript. We would like to thank S. Ida, S. Inutsuka. E. Kokubo and Y. Suto for stimulating conversations and encouragements. This work was supported in part (H.A) by a Japanese Grant-in-Aid for Scientific Research from the Ministry of Education, No. 21540252.
\section{Introduction} Hamilton's principle for mechanical systems with non-holonomic constraints has recently been discussed by Flannery \cite{Flannery}. In particular a variational formulation of the equations of motion of a mechanical system was discussed both for holonomic and non-holonomic constraints. It was shown that while the equations of motion for a system with holonomic constraints can be obtained as variational equations, with the constraints being taken into account by the multiplication rule in the calculus of variations \cite{Mikhlin}, the corresponding procedure with non-holonomic constraints leads to equations which differ from the equations of motion which follow from the well-known principle of d'Alembert. The questions discussed by Flannery are by no means new; they have been discussed in the literature in several papers. For a selection of references other than those given by Flannery, I refer to some of the references contained in a recent paper\footnote{There are some regrettable misprints in this paper. The equations (17), (26) and (34) in Ref. \cite{CCTRJMP} contain the redundant symbols = 0 at the end of the equations.} by Cronstr\"{o}m and Raita \cite{CCTRJMP}. Even though the equations of motion following from the principle of d'Alembert and from the variational action principle with non-holonomic constraints are different in form, there is still the possibility that the equations in question may have the same solutions. It was demonstrated by Pars \cite{Pars}, that this is not the case for a particular example with a specific non-holonomic constraint in a three-dimensional space. As such, this validates the assertion that the solutions to the two different types of equations of motion are different in general, at least in a space of three dimensions. It has only recently been proved, \cite{CCTRJMP}, \cite{CCMN70}, that this is also the case generally in configuration spaces of dimension $N \geq 3$. Below I discuss an improved version of the proofs in \cite{CCTRJMP} and \cite{CCMN70}, separately for $N= 3$ and $N\geq4$. I consider a general autonomous system with a finite number of degrees of freedom, restricted only by reasonable smoothness conditions. For simplicity I consider only the case of one non-holonomic constraint, which is taken to be linear and homogeneous in the generalised velocities of the system. In principle the proof is valid when the configuration space of the system is a fairly general smooth $N$-dimensional manifold, with $N \geq 3$. However, since the argumentation in the proofs of this paper for the most part is local, it is sufficient to use a co-ordinate formulation and consider only one appropriately chosen co-ordinate patch $D_{q}$, with local co-ordinates designated as $q = (q^{1},...,q^{N})$. In the case of two-dimensional systems there is nothing to prove, since every non-holonomic constraint of the kind considered in this paper can be reduced to an equivalent holonomic constraint when $N=2$. The proof given in this paper implies definitely that the standard variational principle, which is valid for unconstrained systems, and which can be generalised to cover the case of holonomic constraints by using the multiplication rule in the calculus of variations, is not in general consistent with the principle of d'Alembert, if one uses the multiplication rule with a genuinely non-holonomic constraint. \section{The variational principles} Consider an autonomous mechanical system with generalised co-ordinates $q = (q^{1},...,q^{N})$, and velocities $\dot{q} = (\dot{q}^{1},...,\dot{q}^{N})$. Throughout this paper I assume the the variables $q = (q^{1},...,q^{N})$ belong to a chosen co-ordinate patch $D_{q}$, which is a contractible domain in configuration space. The kinetic energy of the system is denoted by by $T$, and the generalised applied forces acting on the system are denoted by $Q_{A}, A = 1,...,N$. It will further be assumed that the generalised forces can be expressed in terms of a generalised potential $V(q, \dot{q})$, \begin{equation} \label{GenV} Q_{A} = - \frac{\partial V}{\partial q^{A}} + \frac{d}{dt}\left (\frac{\partial V}{\partial \dot{q}^{A}}\right ), A = 1,...,N. \end{equation} It is appropriate to introduce a Lagrange function $L$ as follows, \begin{equation} \label{lagrangian} L(q, \dot{q}) := T (q, \dot{q}) - V(q, \dot{q}). \end{equation} If the variables $(q^{1},...,q^{N})$ are not constrained in any way, then, as is well known, the equations of motion of the system follow from the following variational principle, \begin{equation} \label{actionprS} \delta S = 0, \end{equation} where \begin{equation} \label{defS} S := \int dt\,L(q, \dot{q}). \end{equation} The equations of motion are the following Euler-Lagrange equations, \begin{equation} \label{ELeq} \frac{d}{dt}\left (\frac{\partial L(q, \dot{q})}{\partial \dot{q}^{A}}\right ) - \frac{\partial L(q, \dot{q})}{\partial q^{A}} = 0 ,\; A = 1, \ldots, N. \end{equation} I then consider a situation in which the variables $(q^{1},...,q^{N})$ are constrained by a condition of the form \begin{equation} \label{holonomic} \Psi(q^{1},...,q^{N}) = C, \end{equation} where $\Psi(q)$ is a smooth (continuously differentiable) function of its variables and $C$ is a constant. The condition (\ref{holonomic}) is a {\em holonomic} constraint. It is well-known that the equations of motion for the system also in this case can be obtained from the variational principle $\delta S = 0$, under the constraint (\ref{holonomic}). Using the multiplication rule in the calculus of variations \cite{Mikhlin} this leads to the following condition, \begin{equation} \label{mprule1} \delta \int \,dt \left [L(q, \dot{q}) + \lambda\, \Psi(q) \right ] = 0, \end{equation} which involves a Lagrange multiplier $\lambda$. The Euler-Lagrange equations following from Eqn. (\ref{mprule1}) are as follows \begin{equation} \label{mpruleEq} \frac{d}{dt}\left (\frac{\partial L(q, \dot{q})}{\partial \dot{q}^{A}}\right ) - \frac{\partial L(q, \dot{q})}{\partial q^{A}} = \lambda\, \frac{\partial \Psi(q)}{\partial q^{A}},\; A = 1, \ldots, N, \end{equation} The equations (\ref{mpruleEq}) constitute a set of $N$ second-order differential equations, which, under appropriate boundary conditions, and together with the constraint (\ref{holonomic}), are supposed to determine the quantities $q^{1}, q^{2}, \ldots, q^{N}$, as well as the Lagrange multiplier $\lambda$. I then finally consider the case of a {\em non-holonomic} constraint, which is taken to be a linear and homogeneous in the generalised velocity variables $\dot{q}^{1},...,\dot{q}^{N}$. The non-holonomic constraint is thus of the following form, \begin{equation} \label{1nonhol} \sum_{A=1}^{N} a_{A}(q)\,\dot{q}^{A} = 0, \end{equation} where the $N$ quantities $a_{A}(q), A = 1, 2, \ldots, N$ are given smooth (continuously differentiable) functions, which transform as covariant vector components under co-ordinate transformations. A metric $g$ is involved in the notation in Eq. (\ref{1nonhol}), which could also be written as follows, \begin{equation} \label{gAB} \sum_{A, B=1}^{N}g_{AB}(q)\,a^{A}(q)\,\dot{q}^{B} = 0. \end{equation} Tensor-indices are thus lowered or raised with the metric tensor $g_{AB}(q)$ and $g^{AB}(q)$, respectively, where \begin{equation} \label{riseg} \sum_{B=1}^{N} g_{AB}(q)\,g^{BC}(q) = g_{A}^{~~C} := \left \{ \begin{array}{ll} 1 & \mbox{if $A = C$}\\ 0 & \mbox{if $A \neq C$,} \end{array} \right . \end{equation} for $A, C = 1, 2, \ldots, N$. It is clear that that not every component of $a_{A}(q)$ can vanish identically in $D_{q}$, since otherwise there would not be any non-holonomic condition at all in the problem. The question is then whether one can use the analogue of the equations (\ref{mprule1}) for a non-holonomic constraint of the form (\ref{1nonhol}), simply by enforcing the constraint (\ref{1nonhol}) by means of the multiplication rule throughout a domain which is supposed to contain all the competing paths in the variational principle. I now prefer to use the special notation \begin{equation} \label{ell0} L_{0}(q, \dot{q}) = T(q, \dot{q}) - V(q, \dot{q}). \end{equation} This special notation is used as a reminder of the fact that the quantity $L_{0}(q,\dot{q})$ that one should be aware of the potential danger involved in using the quantity $L_{0}$ as a Lagrangian in Hamilton's principle when the constraints are genuinely non-holonomic. In the absence of a constraint of the form (\ref{1nonhol}), the quantity $L_{0}(q, \dot{q})$ would naturally be the Lagrangian of the system. The following condition is a straightforward analogue of the variational condition (\ref{mprule1}), \begin{equation} \label{varnh3} \delta \int \,dt \left [L_{0}(q, \dot{q}) - \mu\sum_{A=1}^{N} \,a_{A}(q) \dot{q}^{A} \right ] = 0, \end{equation} which involves a Lagrange multiplier $\mu$. The variational principle underlying the condition (\ref{varnh3}) is, as such, a possible variational principle for certain types of problems, but it does not necessarily give rise to appropriate equations of motion in Newtonian classical mechanics. The condition (\ref{varnh3}) gives rise to the following Euler-Lagrange equations, \begin{equation} \label{CC1} \frac{d}{dt}\left (\frac{\partial L_{0}(q, \dot{q})}{\partial \dot{q}^{A}}\right ) - \frac{\partial L_{0}(q, \dot{q})}{\partial q^{A}} = \dot{ \mu} \,a_{A}(q) + \mu\, \sum_{B =1}^{N} \;M_{AB}(q) \dot{q}^{B},\; A = 1, \ldots, N, \end{equation} where the quantities $M_{AB}(q), A, B = 1. 2. \ldots, N$ are given as follows, \begin{equation} \label{matrixM} M_{AB}(q) := \frac{\partial a_{A}(q)}{\partial q^{B}} - \frac{\partial a_{B}(q)}{\partial q^{A}}, \; A, B = 1, \ldots, N. \end{equation} The variational equations of motion (\ref{CC1}) are precisely those which have been proposed from time to time in the literature, as appropriate variational equations of motion for mechanical systems with non-holonomic constraints of the type considered in this paper. It should be noted that the equations (\ref{CC1}) and the constraint (\ref{1nonhol}), admit a first integral for the system, \begin{equation} \label{energyint} E := \sum_{A=1}^{N}\, \dot{q}^{A}\, \frac{\partial L_{0}(q, \dot{q})}{\partial \dot{q}^{A}} - L_{0}(q, \dot{q}), \end{equation} where $E$ is the constant energy of the system. Consider now the special case in which the constraint (\ref{1nonhol}) is integrable. This means that there exists a function $\Psi(q)$, say, such that \begin{equation} \label{integr} a_{A}(q) = \frac{\partial}{\partial q^{A}} \Psi(q), A = 1, 2, \ldots, N. \end{equation} Then the constraint (\ref{1nonhol}) is in fact a holonomic constraint of the kind given in Eq. (\ref{holonomic}). Inserting the expressions (\ref{integr}) in the equations (\ref{CC1}) one obtains the following equations, \begin{equation} \label{CC1hol} \frac{d}{dt}\left (\frac{\partial L(q, \dot{q})}{\partial \dot{q}^{A}}\right ) - \frac{\partial L(q, \dot{q})}{\partial q^{A}} = \dot{ \mu} \, \frac{\partial \Psi(q)}{\partial q^{A}},\; A = 1, \ldots, N, \end{equation} which naturally agree with the equations (\ref{mpruleEq}) when one makes the change of notation $\dot{\mu} \rightarrow \lambda$. \section{Genuinely non-holonomic constraints} \label{seconstr} It was noted above that if there exists a function $\Psi(q)$ such that the condition (\ref{integr}) holds true, then the constraint (\ref{1nonhol}) is equivalent to a holonomic constraint of the kind given in Eq. (\ref{holonomic}). If the condition (\ref{integr}) holds true, then necessarily, \begin{equation} \label{closeM} M_{AB}(q) = 0, \;\;A, B = 1, 2, \ldots, N, \; q \in D_{q}, \end{equation} where the quantities $M_{AB}, A, B = 1, 2, \ldots, N$ were defined in Eqns. (\ref{matrixM}). Conversely, in a contractible domain, the integrability conditions (\ref{closeM}) are also sufficient for the integrability of the constraint (\ref{1nonhol}). Thus, the constraint (\ref{1nonhol}) is integrable, and therefore equivalent to a holonomic constraint, if and only if the integrability conditions (\ref{closeM}) are fulfilled. In the case at hand, the constraint (\ref{1nonhol}) is equivalent to the following constraint, \begin{equation} \label{intfact} \Phi(q)\, \sum_{A=1}^{N}\,a_{A}(q)\,\dot{q}^{A} = 0, \end{equation} where $\Phi(q)$ is some appropriate smooth function, which does not vanish identically. It may be that one can choose the function $\Phi(q)$ in (\ref{intfact}) in such a manner that this constraint is integrable. If that is the case one says that $\Phi(q)$ is an integrating factor. Replacing $a_{A}$ by $\Phi\,a_{A}, A = 1, 2, \ldots, N$ in the integrability conditions (\ref{closeM}), one finds the following necessary conditions for the existence of an integrating factor, \begin{equation} \label{intcomp} a_{A}(q)M_{BC}(q) + a_{B}(q)M_{CA}(q) + a_{C}(q)M_{AB}(q) = 0, \, A, B, C = 1,2,\ldots, N. \end{equation} The number of independent integrability conditions of the type (\ref{intcomp}) is $N_{c}$, where \cite{Ince}, \begin{equation} \label{numbint} N_{c} = \frac{1}{2}(N-1)(N-2). \end{equation} In three-dimensional space ($N = 3$) there is thus only one integrability condition of the type (\ref{intcomp}), namely the following condition, \begin{equation} \label{3dim} a_{1}(q)M_{23}(q) + a_{2}(q)M_{31}(q) + a_{3}(q)M_{12}(q) = 0. \end{equation} In the case of a contractible domain the integrability conditions (\ref{intcomp}) are also sufficient for the existence of an integrating factor $\Phi(q)$. Thus, if there exists an integrating factor $\Phi(q)$, then the original constraint (\ref{1nonhol}), which has the appearance of a non-holonomic constraint, can be replaced by an equivalent holonomic constraint by making use of the integrating factor. In what follows it is important to consider only such constraints which are {\em genuinely non-holonomic}. A genuinely non-holonomic constraint, linear and homogeneous in the velocity components $\dot{q}^{A}, A = 1, 2, \ldots, A$, is a constraint of the form (\ref{1nonhol}) which is neither integrable as such, nor integrable by means of an integrating factor. Thus, for a genuinely non-holonomic constraint neither the integrability conditions (\ref{closeM}) nor the integrability conditions (\ref{intcomp}) are fulfilled. \section{The d'Alembert equations with constraints} The principle of d'Alembert (see e.g. the classical texts by Goldstein \cite{Goldstein59} or Whittaker \cite{Whittaker}) gives the following equation, \begin{equation} \label{Lag0} \sum_{A=1}^{N} \left \{\frac{d}{dt}\left (\frac{\partial T}{\partial \dot{q}^{A}} \right ) - \frac{\partial T}{\partial q^{A}} - Q_{A}\right \} \delta q^{A} = 0, \end{equation} where the quantities $\delta q^{A}$ are {\em virtual displacements} of the system. The meaning of the symbols $T$ and $Q_{A}$, respectively, is the same as before, {\em i.e.} $T$ is the kinetic energy and $Q_{A}$ is the $A$:th component of the generalised external force imposed on the system. If the virtual displacements $\delta q^{A}, A = 1,\ldots, N$, are independent, then Eq. (\ref{Lag0}) results in the ordinary d'Alembertian equations of motion, \begin{equation} \label{Lag1} \frac{d}{dt}\left (\frac{\partial T}{\partial \dot{q}^{A}} \right ) - \frac{\partial T}{\partial q^{A}} = Q_{A}, \; A = 1, \ldots, N. \end{equation} Consider the generalisation of the discussion above to systems with a constraint of the form (\ref{1nonhol}) given above. The derivation given below of the equations of motion for this system, with a potentially genuinely non-holonomic constraint, can be found in the textbook by Whittaker \cite{Whittaker}. Implement the constraint (\ref{1nonhol}) by regarding the system to be acted on by both the external applied forces $Q_{A}, \, A=1,\ldots,N$, and by certain additional forces of constraint $Q'_{A},\, A=1,\ldots,N$, which force the system to satisfy the condition (\ref{1nonhol}). Equation~(\ref{Lag0}) is then replaced by the following equation, \begin{equation} \label{Lag3} \sum_{A=1}^{N}\,\left \{\frac{d}{dt}\left (\frac{\partial T}{\partial \dot{q}^{A}} \right ) - \frac{\partial T}{\partial q^{A}} - Q_{A} - Q'_{A} \right \} \delta q^{A} = 0, \end{equation} In Eq. (\ref{Lag3}) the virtual displacements $\delta q^{A}, A = 1, \ldots, N$, can now be regarded as independent. Thus one obtains the equations of motion, \begin{equation} \label{Lagnh} \frac{d}{dt}\left (\frac{\partial T}{\partial \dot{q}^{A}} \right ) - \frac{\partial T}{\partial q^{A}} =Q_{A} + Q'_{A}, \,A = 1, \ldots, N. \end{equation} The forces of constraint, $Q'_{A}, A = 1, \ldots, N$, are {\em a priori} unknown, but they are such that, in any instantaneous displacement $\delta q^{A}, A = 1,\ldots, N$, consistent with the constraint (\ref{1nonhol}), they do no work. The non-holonomic constraint (\ref{1nonhol}) implies the following condition on the possible instantaneous displacements $\delta q^{A}, A = 1,\ldots, N$, \begin{equation} \label{cdelta} \sum_{A=1}^{N}\,a_{A}(q) \delta q^{A} = 0. \end{equation} For any instantaneous displacements $\delta q^{A}, A = 1,\ldots, N$, which satisfy the condition (\ref{cdelta}), the work $\delta W'$ done by the constraint forces $Q'_{A}, A=1,\ldots,N$ must be equal to zero, {\em i.e.}, \begin{equation} \label{workc} \delta W' := \sum_{A=1}^{N} Q'_{A}\delta q^{A} = 0. \end{equation} The conditions (\ref{cdelta}) and (\ref{workc}) together imply that \begin{equation} \label{cforce} Q'_{A} = \lambda a_{A}(q),\, A = 1,\ldots, N, \end{equation} where the quantity $\lambda$ is a time-dependent parameter to be determined. Eqns. (\ref{Lagnh}) have thus been reduced to the following form, \begin{equation} \label{Lagnh2} \frac{d}{dt}\left (\frac{\partial T}{\partial \dot{q}^{A}} \right ) - \frac{\partial T}{\partial q^{A}} =Q_{A} + \lambda a_{A}(q), \,A = 1, \ldots, N. \end{equation} These $N$ equations of motion are consequences of the principle of d'Alembert. They will be called {\em d'Alembertian equations}. One should still add the equation of constraint (\ref{1nonhol}) to the equations of motion above. There are thus altogether $N+ 1$ equations for the determination of $N+1$ quantities $q^{A}(t), A = 1,\ldots,N$, and $\lambda(t)$, when appropriate boundary conditions for the quantities $q^{1},...,q^{N}$ and $\dot{q}^{1},...,\dot{q}^{N}$ are given. One should observe that in the argument above, it is \underline{not} required that the constraint equation (\ref{1nonhol}) be in force under general variations $q_{j} \rightarrow q_{j} + \delta q_{j}$; the constraint (\ref{1nonhol}) is \underline{only} imposed on the actual motion of the system. Assuming that the generalised force components $Q_{A}, A = 1, 2, \ldots, N$ can be expressed in terms of a generalised potential $V(q, \dot{q})$, as in Eq. (\ref{GenV}), one can rewrite the d'Alembertian equations (\ref{Lagnh2}) as follows, \begin{equation} \label{Lagnh3} \frac{d}{dt}\left (\frac{\partial L_{0}(q, \dot{q})}{\partial \dot{q}^{A}}\right ) - \frac{\partial L_{0}(q, \dot{q})}{\partial q^{A}} = \lambda a_{A}(q), \;\,A = 1, \ldots, N, \end{equation} where the quantity $L_{0}(q, \dot{q})$ has been defined in Eq. (\ref{ell0}). The equations (\ref{Lagnh3}) will also be referred to as d'Alembertian equations in what follows. The d'Alembertian equations of motion (\ref{Lagnh3}) admit a first integral, namely the expression (\ref{energyint}). This is the same first integral as the one obtained from the variational equations (\ref{CC1}). \section{Non-equivalence of the principle of d'Alembert and the variational action principle with genuinely non-holonomic constraints} I now consider the d'Alembertian equations of motion (\ref{Lagnh3}) and the variational equations (\ref{CC1}), respectively, as {\em initial value} problems. This means that the co-ordinate patch in which the quantities $q^{1}, q^{2}, \ldots, q^{N}$ are local co-ordinates, is assumed to be chosen so that it includes the initial value point $q_{0} = (q_{0}^{1}, q_{0}^{2}, \ldots, q_{0}^{N})$. The question is whether the equations (\ref{Lagnh3}) and (\ref{CC1}), respectively, can have the same solutions for $q^{1}(t), \ldots, q^{N}(t)$ in general, despite the fact that these equations are not identical. It will be shown below that this is not the case. Naturally, the functions $L_{0}(q, \dot{q})$ and $a_{A}(q)$ entering into the equations (\ref{Lagnh3}) and (\ref{CC1}), respectively, will have to satisfy appropriate smoothness conditions in order that these equations may have solutions. I will not enter into a discussion of such smoothness conditions, but rather assume that the equations (\ref{Lagnh3}) and (\ref{CC1}), respectively, have {\em e.g.} $C^{2}$-solutions in some appropriate time-interval, for given initial values for the co-ordinates $q^{A}$ in the domain $D_{q}$ and for given initial values for the velocities $\dot{q}^{A}, A = 1, \ldots, N$, \begin{equation} \label{initq} \left [q^{A}(t)\right ]_{t=t_{0}} = q_{0}^{A},\; \left [\dot{q}^{A}(t)\right ]_{t=t_{0}} = \dot{q}_{0}^{A},\; A = 1, \ldots, N. \end{equation} The initial values $ q_{0}^{A}$ and $\dot{q}_{0}^{A}$ at $t = t_{0}$ are free parameters within an appropriate region of the configuration- and velocity space, except for the restriction \begin{equation} \label{initrestr} \sum_{A=1}^{N} \,a_{A}(q_{0})\, \dot{q}_{0}^{A} = 0. \end{equation} The condition (\ref{initrestr}) is a consequence of the non-holonomic constraint (\ref{1nonhol}). After these considerations concerning the initial values I will prove a theorem on the incompatibility of the d'Alembertian and variational equations of motion, respectively. \subsection{The incompatibility theorem} \label{incomp} Consider the d'Alembertian and variational equations of motion (\ref{Lagnh3}) and (\ref{CC1}), respectively, which involve the genuinely non-holonomic constraint (\ref{1nonhol}). If $N=3$ the d'Alembertian and variational equations of motion are not compatible in the sense that they do not have coinciding solutions. If $N\geq4$ the d'Alembertian and variational equations of motion are not compatible in the sense that they do not have coinciding solutions with arbitrary general initial values (\ref{initq}), which satisfy the condition (\ref{initrestr}). It is assumed that the d'Alembertian equations (\ref{Lagnh3}) have unique smooth ({\em e.g.} $C^{2}$) solutions $(q^{1}(t), \ldots, q^{N}(t))$ in an appropriate time-interval, satisfying a genuinely non-holonomic constraint of the type (\ref{1nonhol}), and the general initial value conditions (\ref{initq}) and (\ref{initrestr}). I call such solutions general solutions in what follows. The method of proof is by {\em reductio ad absurdum}, {\em i.e.}, one makes the assumption that the variational equations (\ref{CC1}) have solutions which coincide with the general solutions of the d'Alembertian equations (\ref{Lagnh3}), and shows that this assumption leads to contradictions. The proof given below is designed in such a manner that it is independent of any specific properties of the kinetic energy $T$ and generalised potential $V$, respectively, for the problem under consideration. \subsection{Preliminary considerations} \label{preliminary} Assume now that the equations (\ref{CC1}) and (\ref{Lagnh3}) have coincident solutions. By subtracting Eqns. (\ref{CC1}) from Eqns. (\ref{Lagnh3}), one obtains the following equations, \begin{equation} \label{diffdACC1} (\lambda - \dot{\mu})\,a_{A}(q) = \mu \; \sum_{B=1}^{N} M_{AB}(q)\,\dot{q}^{B} = 0, \; A = 1, \ldots, N, \end{equation} which have to be satisfied by the general solutions $(q^{1}, \ldots, q^{N})$ of the d'Alembertian equations (\ref{Lagnh3}). It should be noted that one must necessarily have $\mu \not\equiv 0$ in the equations (\ref{diffdACC1}) above, since otherwise the variational equations (\ref{CC1}) would contain no reference whatsoever to the constraints (\ref{1nonhol}). Further, if $\mu \equiv 0$, it follows from the equations (\ref{diffdACC1}) that also the following conditions must hold true, \begin{equation} \label{exceptA} \lambda\,a_{A} \equiv 0,\; A = 1. \ldots, N. \end{equation} The conditions (\ref{exceptA}) above imply that either \begin{equation} \label{lamndanoll} \lambda \equiv 0, \end{equation} or \begin{equation} \label{anoll} a_{A}(q) \equiv 0,\; A = 1. \ldots, N. \end{equation} If the conditions (\ref{anoll}) are true, there are no constraints to be considered so the the basic question under consideration disappears. Finally, if the condition (\ref{lamndanoll}) holds true in addition to the condition $\mu \equiv 0$, then the d'Alembertian equations of motion (\ref{Lagnh3}) and the variational equations of motion (\ref{CC1}) are trivially identical, so the the basic question under consideration disappears also in that case. Hence one must have $\mu \not\equiv 0$. Using the notation \begin{equation} \label{Gamma} \Gamma := \frac{\lambda - \dot{\mu}}{\mu}. \end{equation} \noindent one can write the conditions (\ref{diffdACC1}) in an equivalent form as follows, \begin{equation} \label{eqnsM} \sum_{B=1}^{N} M_{AB}(q)\,\dot{q}^{B} = \Gamma\,a_{A}(q), \; A = 1, \ldots, N, \end{equation} The tensor equations (\ref{eqnsM}) constitute a set of $N$ linear algebraic equations in the variables $\dot{q}^{A}, A = 1, \ldots, N$, for any fixed value of $q := (q^{1}, q^{2}, \ldots, q^{N}) \in D_{q}$. It should be noted that the quantities $M_{AB}(q)$ in Eqns. (\ref{eqnsM}) are anti-symmetric upon an interchange $A \leftrightarrow B$ of the indices $A$ and $B$, \begin{equation} \label{antisym} M_{AB}(q) = - M_{BA}(q), \; A, B = 1, 2, \ldots, N. \end{equation} Thus the matrix-like quantity $M(q) := (M_{AB}(q))$ is {\em skew}. \subsection{The case $N = 3$} \label{dimension3} I consider the case N = 3 separately, since in that case the analysis of the tensor equations (\ref{eqnsM}) is particularly simple. Thus, for $N = 3$ the equations (\ref{eqnsM}) read as follows, \begin{equation} \label{eqnsM3} \sum_{B=1}^{3} M_{AB}(q)\,\dot{q}^{B} = \Gamma\,a_{A}(q), \; A = 1, 2, 3. \end{equation} It should be noted that since $N = 3$ and $M(q)$ is skew, one necessarily has \begin{equation} \label{detM3} \det M(q) = 0. \end{equation} The quantity $\Gamma$ in Eqns. (\ref{eqnsM3}) is a so far unknown parameter, but one must have either $\Gamma \equiv 0$ or $\Gamma \not\equiv 0$, Consider first the case $\Gamma \equiv 0$. Then the velocity components $(\dot{q}^{1}, \dot{q}^{2}, \dot{q}^{3})$ satisfy the homogeneous equations \begin{equation} \label{homM3} \sum_{B=1}^{3} M_{AB}(q)\,h^{B}(q) = 0, \; A = 1, 2, 3. \end{equation} The homogeneous equations (\ref{homM3}) have the following non-trivial solutions, \begin{equation} \label{solM3} (h^{1}(q), h^{2}(q), h^{3}(q)) := \alpha ( M_{23}(q), M_{31}(q), M_{12}(q)), \; \alpha \not\equiv 0. \end{equation} However, the solutions (\ref{solM3}) for $(\dot{q}^{1}, \dot{q}^{2}, \dot{q}^{3})$ must also satisfy the constraint (\ref{1nonhol}) with $N=3$, {\em i.e.}, \begin{equation} \label{const3} \sum_{A=1}^{N} a_{A}(q)\, \dot{q}^{A} = 0. \end{equation} This constraint then implies that \begin{equation} \label{acond} a_{1}(q)M_{23}(q) + a_{2}(q)M_{31}(q) + a_{3}(q)M_{12}(q) = 0. \end{equation} The equation (\ref{acond}) is nothing but the condition (\ref{3dim}), which implies that the constraint (\ref{const3}) is integrable by means of an integrating factor. This is a contradiction, since by assumption the constraint (\ref{const3}) is genuinely non-holonomic. If $\Gamma \not\equiv 0$, the equations (\ref{eqnsM3}) are inhomogeneous equations. In order that the inhomogeneous equations (\ref{eqnsM3}) may have a solution it is necessary that the quantity $a = (a_{1}, a_{2}. a_{3})$ on its right hand side be orthogonal to any non-trivial solution (\ref{solM3}) of the homogeneous equation (\ref{homM3}). This also leads to the condition (\ref{acond}) as a necessary condition for the solvability of Eqns. (\ref{eqnsM3}). This implies the same contradiction as in the homogeneous case. It has thus been proved that the d'Alembertian equations of motion (\ref{Lagnh3}) and the variational equations of motion (\ref{CC1}) in the case $N= 3$ can not have coincident solutions if the constraint (\ref{const3}) is genuinely non-holonomic. \subsection{The cases $N \geq 4$} \label{dimens4} I will now consider the linear equations (\ref{eqnsM}) for any finite dimension $N \geq 4$, at some fixed point $q = (q^{1}, q^{2}, \ldots, q^{N}) \in D_{q}$. The equations (\ref{eqnsM}), which are necessary consequences of the assumption that d'Alembertian equations of motion (\ref{Lagnh3}) and the variational equations of motion (\ref{CC1}) have coincident solutions, are valid at any given fixed point $q \in D_{q}$. It is now appropriate to analyse the equations (\ref{eqnsM}) in more detail. Contract Eqns (\ref{eqnsM}) with the quantity $(a^{1}(q), a^{2}(q), \ldots, a^{N}(q))$. This leads to the result \begin{equation} \label{GamtoR} \sum_{A, B=1}^{N} a^{A}(q) M_{AB}(q) \dot{q}^{B} = \Gamma ||a(q)||^{2}, \end{equation} where the norm $||a(q)||$ is defined by the expression \begin{equation} \label{norma} ||a(q)||^{2} = \sum_{A=1}^{N} a^{A}(q)a_{A}(q). \end{equation} Let $R(q, \dot{q})$ be defined as follows, \begin{equation} \label{Def1R} R(q, \dot{q}) := \sum_{A, B=1}^{N} a^{A}(q)M_{AB}(q)\,\dot{q}^{B}. \end{equation} Trading the quantity $\Gamma$ for $R(q, \dot{q})$ in Eqns. (\ref{eqnsM}) one obtains, \begin{equation} \label{eqnsR} \sum_{D=1} M^{A}_{~~D}(q)\dot{q}^{D} = \frac{a^{A}(q) R(q, \dot{q})}{||a(q)||^{2}||}, A = 1, 2, \ldots, N, \end{equation} where the free index $A$ is raised for convenience. Contracting Eq. (\ref{eqnsR}) with the quantity $\sum_{B=1}^{N}M_{AB}\dot{q}^{B}$, and using the anti-symmetry condition (\ref{antisym}), one gets finally the following expression for the quantity $R(q, \dot{q})$ defined in Eq, (\ref{Def1R}), \begin{equation} \label{Rsquare} R^{2}(q, \dot{q}) = - ||a(q)||^{2}\, \sum_{B, D=1}^{N} (M^{2})_{BD}(q)\dot{q}^{B}\dot{q}^{D}, \end{equation} where \begin{equation} \label{Msquare} (M^{2})_{BD}(q) := \sum_{A=1}^{N} M_{BA}(q)M^{A}_{~~D}(q), \; B, D = 1, 2, \ldots, N. \end{equation} From the anti-symmetry condition (\ref{antisym}) follows also that the quantity $M^{2}$ is symmetric in its lower indices, \begin{equation} \label{Msqsymm} (M^{2})_{BD}(q) = (M^{2})_{DB}(q), \; B, D = 1, 2, \ldots, N. \end{equation} Incidentally, Eq. (\ref{Rsquare}) shows that the following quadratic form $Q_{(2)}(x)$ in real variables $x^{A}, A = 1, 2, \ldots, N$, is negative semi-definite, \begin{equation} \label{defQ2} Q_{(2)}(x) ;= \sum_{B, D=1}^{N} \sum_{A=1}^{N} M_{BA}(q)M^{A}_{~~D}(q)x^{B}x^{D} \leq 0. \end{equation} This is made even more explicit by a reshuffling of the summation indices in Eq. (\ref{defQ2}), \begin{equation} \label{innerpQ} Q_{(2)}(x) = - \sum_{A, C=1}^{N} g^{AC}(q) \left (\sum_{B=1}^{N} M_{AB}(q)x^{B} \right ) \left (\sum_{D=1}^{N} M_{CD})q) x^{D} \right ) \equiv - ||M(q)x||^{2}. \end{equation} Thus the quadratic form $Q_{(2)}(x)$ vanishes if and only if \begin{equation} \label{nullQ2} \sum_{B=1}^{N} M_{AB}(q)x^{B} = 0. \end{equation} The results given above will be used subsequently. Applying the formula (\ref{innerpQ}) to the expression (\ref{Rsquare}) one gets finally the following useful expression, \begin{equation} \label{Rfinale} R(q, \dot{q}) = + ||a(q)|| \, ||M(q)\dot{q}||. \end{equation} There is an ambiguity involving a sign $\pm 1$ in arriving at the expression (\ref{Rfinale}) along the route indicated above. However this sign-ambiguity is of no consequence for the remaining considerations in this paper, so for simplicity I shall stick to the sign-factor $+ 1$ in the expression (\ref{Rfinale}) in what follows. Inserting the expression (\ref{Rfinale}) for $R(q, \dot{q})$ in Eqns. (\ref{eqnsR}) one finds the following equations, \begin{equation} \label{eqnsRfinal} \sum_{B=1}^{N} M_{AB}(q)\dot{q}^{B} = \frac{||M(q)\dot{q}||}{||a(q)||}\, a_{A}(q), A = 1, 2, \ldots, N. \end{equation} The tensor equations (\ref{eqnsRfinal}) are more precise versions of the equations (\ref{eqnsM}), which constituted the starting point in the proof of the non-compatibility theorem. It is necessary to consider the detailed structure of the solutions $\dot{q}^{A}, A = 1, 2, \ldots, N$, of Eqns. (\ref{eqnsRfinal}) in order to proceed with the proof of the non-compatibility theorem enunciated in Subsection \ref{incomp}. For this purpose one needs certain basic facts concerning skew matrix-like quantities such as the anti-symmetric tensor fields $(M_{AB}(q))$ in the equations above. \subsection{Digression on skew matrices $(M_{AB})$} \label{digrM} In this subsection I will discuss such properties related to the skew matrix-like quantities $M_{AB}(q), \,A, B = 1, 2, \ldots, N$, encountered previously, which are needed for finding solutions of Eqns. (\ref{eqnsRfinal}) for $\dot{q}^{A}, A = 1, 2, \ldots, N$. The components $g_{AB}(q), \; A, B = 1, 2, \ldots, N$ of the metric tensor $g$ enter in an essential way in this discussion. The co-ordinates $q$ enter as parameters in the discussion below, which deals with exclusively algebraic properties. In this discussion the co-ordinates $q$ may either considered to correspond to a fixed point $q_{0} \in D_{q}$, or else to be confined to an appropriate reducible sub-domain $D^{0}_{q} \subset D_{q}$, which contains the given fixed point $q_{0}$. Eventually the fixed point $q_{0}$ will be identified with an initial value in co-ordinate space for the d'Alembertian and variational equations of motion under consideration in this paper. Consider the following quadratic form $G_{(2)}(x)$ in certain real variables $x^{A}, A = 1, 2, \ldots, N$, \begin{equation} \label{G2} G_{(2)}(x) := \sum_{A, B=1}^{N} g_{AB}(q)x^{A}x^{B}. \end{equation} It follows from the assumed properties of the metric $g(q)$, that the quadratic form $G_{(2)}(x)$ is a real positive definite quadratic form. I also recall the negative semi-definite quadratic form $Q_{(2)}(x)$ defined above in Eq. (\ref{defQ2}), involving the quantities $(M^{2})_{BD} = (M^{2})_{DB}, \, B, D = 1, 2, \ldots, N$ defined in Eq. (\ref{Msquare}). It is known that one can effect a simultaneous reduction of two quadratic forms, of which one is a positive definite form such as the form (\ref{G2}) above, and the other is merely a real quadratic form, such as the form (\ref{defQ2}) above, to sums of perfect squares \cite{Friedman}. This leads to the following eigen-value problem for eigen-vectors $e^{A}(q), A = 1, 2, \ldots, N$, \begin{equation} \label{eigenv} \sum_{B=1}^{N} \left ( (M^{2})_{AB}(q) - \lambda(q) \, g_{AB}(q) \right )e^{B}(q) = 0. \end{equation} The equation (\ref{eigenv}) has non-trivial solutions only if the parameter $\lambda$ satisfies the following eigen-value equation, which is similar to the characteristic equation for eigen-values in ordinary linear algebra, \begin{equation} \label{detCC1} \det \left ( (M^{2})_{AB}(q) - \lambda(q)\, g_{AB}(q) \right )= 0. \end{equation} Under the conditions described above for the quadratic forms $G_{(2)}(x)$ given in (\ref{G2}), and the quadratic form $Q_{(2)}(x)$ given in (\ref{defQ2}), the equation (\ref{detCC1}) has $N$ eigen-value solutions $\lambda_{\nu}(q), \nu = 1, 2, \ldots, N$. The corresponding eigen-vectors $e_{\nu}^{~B}(q)$ satisfy the following equations, \begin{equation} \label{eignu} \sum_{B=1}^{N} \left ( (M^{2})_{AB}(q) - \lambda_{\nu}(q) \, g_{AB}(q) \right )e_{\nu}^{~B}(q) = 0, \: \;\nu = 1, 2, \ldots, N. \end{equation} It is convenient to consider the following variant of Eqns. (\ref{eignu}), \begin{equation} \label{eignulow} \sum_{B=1}^{N} \left ( (M^{2})_{A}^{~~B}(q) - \lambda_{\nu}(q) \, g_{A}^{~~B}(q) \right )e_{\nu B}(q) = 0, \;\;\nu = 1, 2, \ldots, N. \end{equation} The equations (\ref{eignu}) and (\ref{eignulow}) are perfectly equivalent; the equations (\ref{eignulow}) are obtained by simultaneously raising and lowering the summation indices $B$ inside the equations (\ref{eignu}), which is a straightforward and legitimate operation. From Eqns. (\ref{eignulow}) follows that the eigen-values $\lambda_{\nu}(q), \nu = 1, 2, \ldots, N$, also satisfy the following variant of Eq. (\ref{detCC1}), \begin{equation} \label{detCC2} \det \left ( (M^{2})_{A}^{~~B}(q) - \lambda_{\nu}(q) \, g_{A}^{~~B}(q) \right )= 0. \end{equation} Here the enumeration of rows and columns in the determinant is as follows: the lower indices $A = 1, 2, \ldots, N$ enumerate the rows, and the upper indices $B = 1, 2, \ldots, N$ enumerate the columns. The equation (\ref{detCC2}) has an advantage over Eq. (\ref{detCC1}), namely that the metric $g$ is only involved in the expressions $ (M^{2})_{A}^{~~B}(q), A, B = 1, 2, \ldots, N$, in Eq. (\ref{detCC2}). This is due to the fact that the quantities $g_{A}^{~~B}(q), A, B = 1,2, \ldots, N$ always have the fixed numerical values $0$ or $1$, independently of the co-ordinates $q$, \begin{equation} \label{gdelta} g_{A}^{~~B}(q) := \left \{ \begin{array}{ll} 1 & \mbox{if $A = B$}\\ 0 & \mbox{if $A \neq B$,} \end{array} \right . \end{equation} for $A, B = 1, 2, \ldots, N$. It should also be noticed that it follows readily from Eq. (\ref{detCC2}) that the eigen-values $\lambda_{\nu}(q), \nu = 1, 2, \ldots, N$, are scalars, {\em i.e.} invariant under co-ordinate transformations. I finally rewrite Eqns. (\ref{eignulow}) as follows, \begin{equation} \label{finaleigen} \sum_{B=1}^{N} (M^{2})_{A}^{~~B}(q)\, e_{\nu B}(q) = \lambda_{\nu}(q)\, e_{\nu A}(q), \; \nu = 1, 2, \ldots, N, \; A = 1, 2, \ldots, N. \end{equation} It should be noted that the eigen-vectors $e_{\nu A}(q), \; A = 1, 2, \ldots, N$ are orthogonal vectors, as will be shown below. If the eigen-values $\lambda_{\nu}(q), \nu = 1, 2, \ldots, N$ are non-degenerate, this follows immediately from a consideration the equation (\ref{finaleigen}) for two different indices $\mu$ and $\nu$, say. Contract the equations (\ref{finaleigen}) with $e_{\mu}^{~A}(q)$. Interchange $\mu$ and $\nu$ in the equations obtained in this way, and subtract one set of equations from the other set of equations in the pairs of equations obtained by the contraction procedure. Then one obtains, \begin{equation} \label{ortho} (\lambda_{\mu}(q) - \lambda_{\nu}(q))(e_{\mu A}(q), e_{\nu B}(q)) = 0, \end{equation} where \begin{equation} \label{ginnerprod} (e_{\mu}(q), e_{\nu}(q)) := \sum_{A, B=1}^{N} g^{AB}(q)e_{\mu A}(q)e_{\nu B}(q). \end{equation} The abbreviated notation on the left hand side of the expression (\ref{ginnerprod}) for the inner product of vector-quantities such as the eigen-vectors $e_{\nu A}(q), \; A = 1, 2, \ldots, N$, will be frequently used in what follows. If the eigen-values are non-degenerate, i.e., if $\lambda_{\mu}(q) \neq \lambda_{\nu}(q)$ when $\mu \neq \nu$, then the orthogonality of the corresponding eigen-vectors $e_{\mu}(q)$ and $e_{\nu}(q)$ follows from the equations (\ref{ortho}). Normalising the eigen-vectors to $1$ in the inner product (\ref{ginnerprod}), one then finally obtains the following ortho-normalisation conditions, \begin{equation} \label{orthonorm} (e_{\mu}(q), e_{\nu}(q)) = \delta_{\mu\nu} := \left \{ \begin{array}{ll} 1 & \mbox{if $\mu = \nu$}\\ 0 & \mbox{if $\mu \neq \nu$,} \end{array} \right . \end{equation} for $\mu, \nu = 1, 2, \ldots, N$. If there are degenerate eigen-values, the ortho-normalisation conditions are imposed as a part of the definition of the corresponding eigen-vectors. Thus the ortho-normalisation conditions (\ref{orthonorm}) are valid in all cases. Using the ortho-normalisation conditions (\ref{orthonorm}) one readily obtains the eigen-values $\lambda_{\nu}(q), \nu = 1, 2, \ldots, N$, from Eqns. (\ref{finaleigen}), \begin{equation} \label{lambdavalue} \lambda_{\nu}(q) = \sum_{A, B=1}^{N} e_{\nu}^{~A}\,(M^{2})_{A}^{~~B}(q)\, e_{\nu B}(q)) \equiv - ||M(q) e_{\nu}(q)||^{2}, \nu = 1, 2, \ldots, N. \end{equation} In arriving at the final expressions above for the eigen-values $\lambda_{\nu}, \nu = 1, 2, \ldots, N$, use has been made of the result given in Eq. (\ref{innerpQ}). From the expressions (\ref{lambdavalue}) follows that all the eigen-values $\lambda_{\mu}, \mu = 1, 2, \ldots, N$ are non-positive, \begin{equation} \label{kappa1} \lambda_{\nu}(q) := - \kappa_{\nu}(q)^{2}, \end{equation} where the real quantities $\kappa_{\nu}(q), \nu = 1, 2, \ldots,, N$, can be chosen to be non-negative without loss of generality. Comparing Eqns. (\ref{lambdavalue}) and (\ref{kappa1}) one then obtains, \begin{equation} \label{kappafin} \kappa_{\nu}(q) = ||M(q) e_{\nu}(q)||,\;\; \nu = 1, 2, \ldots, N. \end{equation} From Eq. (\ref{kappafin}) one gets an upper bound on the quantities $\kappa_{\nu}, \nu = 1, 2, \ldots, N$, in terms of the norm $||M(q)||$ of the matrix $M(q)$, \begin{equation} \label{normM} ||M(q)|| = sup_{||x||=1} ||M(q)\,x||. \end{equation} The quantities $\kappa_{\nu}, \nu = 1, 2, \ldots, N$, thus satisfy the following inequalities, \begin{equation} \label{kappabounds} 0 \leq \kappa_{\nu}(q) \leq ||M(q)||, \; \nu = 1, 2, \ldots, N. \end{equation} From Eq. (\ref{lambdavalue}) follows that an eigen-value $\lambda_{\nu}(q)$ equals zero if and only if \begin{equation} \label{kerM1} \sum_{B=1}^{N} M_{AB}(q)\, e_{\nu}^{~B}(q) = 0, \;\; A = 1, 2, \ldots, N. \end{equation} The number of non-vanishing eigen-values is thus determined by the {\em rank} of the matrix $M(q)$. Since $M(q)$ is skew, the rank of $M(q)$ is necessarily an even integer, $2p$, say, where $p \geq 1$, since $M(q) \not\equiv 0$. The eigen-values will now be enumerated in such a manner that the last $N-2p$ eigen-values $\lambda_{\nu}, \nu = 2p+1, \ldots, N$ are equal to zero. Thus, \begin{equation} \label{kerM} \sum_{B=1}^{N} M_{AB}(q)\, e_{\nu}^{~B}(q) = 0, \;\;A = 1, 2, \ldots, N;\, \nu = 2p+1, \ldots, N. \end{equation} The subspace $ker(M(q))$ is thus spanned by the eigen-vectors $e_{\nu}(q), \, \nu = 2p+1, \ldots, N$, the components of which have been labelled by upper or indices $A = 1, 2, \ldots, N$, in Eqns. (\ref{kerM}). One may also use lower indices $A, B,\ldots$, whenever it is convenient, by using the standard rules for lowering or raising such indices with the metric tensor $g_{AB}(q)$ or $g^{AB}(q)$, respectively, where $A, B = 1, 2, \ldots, N$. Corresponding to the enumeration in Eqns. (\ref{kerM}) one recognises that \begin{equation} \label{nullkappa} \kappa_{\nu}(q) = 0, \; \nu = 2p+1, \ldots, N. \end{equation} If $M(q)$ is regular there are no non-trivial solutions of the equations (\ref{kerM}). This can happen only if $N$ is even and when $2p = N$. Generally, for $\nu = 1, \ldots, 2p$, the quantities $\kappa_{\nu}(q)$ are positive, \begin{equation} \label{poskappa} \kappa_{\nu}(q) > 0, \;\; \nu = 1, 2, \ldots, 2p. \end{equation} A remark on the classification of a skew matrix $M(q)$ according to its rank may be in order. Since $M(q)$ depends on $q$, also the rank $2p$ is in principle dependent on $q$. If considerations related to a skew matrix $M(q)$ involving only one prescribed fixed point $q_{0}$, say, no problems can occur by the omission of the argument $q$ in the quantity $p$. This notation is used here for the sake of simplicity. Strictly speaking, however, one should use the notation $p(q)$ in stead of only $p$. The rank $2p(q)$ takes values in the set of even positive integers, and therefore the rank $2p(q)$ ought not to change from its value at a prescribed point $q_{0}$ if one considers neighbouring points $q$ sufficiently near the prescribed point $q_{0}$. This also justifies the omission of the argument $q$ from the symbol $2p$, which then can be considered to be a valid characterisation of $M(q)$ at least in an appropriate sufficiently small domain containing the prescribed point $q_{0}$ as an interior point. I will now construct a new basis in the $N$-dimensional linear space spanned by the eigen-vectors $e_{\nu}(q), \,\nu = 1, 2, \ldots, N$. This basis is used to solve the important equations (\ref{eqnsRfinal}). The construction parallels closely a similar construction made by Greub \cite{Greub} for skew mappings in ordinary Euclidean space. Define vectors $b_{2\nu-1}(q), \, \nu = 1, \ldots, p$, as follows, \begin{equation} \label{Defbodd} b_{2\nu-1}^{~~~~~A}(q) := e_{\nu}^{~A}(q), \, \nu = 1, \ldots, p\; ; A = 1, 2, \ldots, N. \end{equation} Similarly, let vectors $b_{2\nu}(q), \, \nu = 1, \ldots, p$, be defined by the equations, \begin{equation} \label{Defeven} b_{2\nu A}(q) := \kappa_{\nu}(q)^{-1}\sum_{B=1}^{N}\, M_{AB}(q)b_{2\nu-1}^{~~~~~B}(q), \, \nu = 1, \ldots, p\; ; A = 1, 2, \ldots, N. \end{equation} In view of the conditions (\ref{poskappa}) the definition is a possible definition. I rewrite the eigen-value equations (\ref{finaleigen}) with $- \kappa_{\nu}(q)^{2}$ in stead of $\lambda_{\nu}(q)$, thus \begin{equation} \label{finalkap} \sum_{B=1}^{N} (M^{2})_{A}^{~~B}(q)\, e_{\nu B}(q) = - \kappa_{\nu}(q)^{2}\,(q)\, e_{\nu A}(q), \; \nu = 1, 2, \ldots, N, \; A = 1, 2, \ldots, N. \end{equation} It is now a simple matter to show that \begin{equation} \label{beigen} \sum_{B=1}^{N} (M^{2})_{A}^{~~B}(q)\, b_{\mu B}(q) = - \kappa_{\nu}(q)^{2} \, b_{\mu A}(q), \; \mu = 1, 2, \ldots, 2p, \; A = 1, 2, \ldots, N. \end{equation} When $2p < N$ one also defines, \begin{equation} \label{kerM3} b_{\nu A}(q) := e_{\nu A}(q), \; \nu = 2p+1, \ldots, N,\; , A = 1, 2, \ldots, N. \end{equation} Then, \begin{equation} \label{nulleigen} \sum_{B=1}^{N} (M^{2})_{A}^{~~B}(q)\, b_{\nu B}(q) = 0, \; \nu = 2p+1, \ldots, N \; ;A = 1, 2, \ldots, N. \end{equation} From the definition (\ref{Defeven}) follows that \begin{equation} \label{Meqodd} \sum_{B=1}^{N}\, M_{AB}(q)b_{2\nu-1}^{~~~~~B}(q) = \kappa_{\nu}(q) b_{2\nu A}(q), \;, \nu = 1, \ldots, p\; : \; A = 1, 2, \ldots, N. \end{equation} Using the equations (\ref{beigen}) it is also simple to show that \begin{equation} \label{Meqeven} \sum_{B=1}^{N}\, M_{AB}(q) b_{2\nu}^{~~~B}(q) = - \kappa_{\nu}(q) b_{2\nu-1 A}(q), \;, \nu = 1, \ldots, p\; : \; A = 1, 2, \ldots, N. \end{equation} It is finally stated without proof, that the basis vectors $b_{\mu}(q), \mu = 1, 2, \ldots, N$, defined above, are orthonormal in the inner product (\ref{ginnerprod}), \begin{equation} \label{bortho} (b_{\mu}(q), b_{\nu}(q)) = \delta_{\mu \nu}, \; \mu, \nu = 1, 2, \ldots, N. \end{equation} The proof of the ortho-normalisation conditions (\ref{bortho}) is fairly simple, and is left to the interested reader. The vectors $b_{\mu},\, \mu = 1, 2, \ldots, N$ thus also form an orthonormal basis in the space spanned by the eigen-vectors $e_{\mu}, \, \mu = 1, 2, \ldots, N$ of the matrix $M^{2}(q)$. Incidentally, the results given above imply that there are only $p$ non-zero eigen-values $\lambda_{\nu}(q) = -\kappa_{\nu}(q)^{2}$ related to the matrix $M^{2}(q)$ in stead of $2p$ such values. This means that the non-zero eigen-values of $M^{2}(q)$ are at least two-fold degenerate. \subsection{The solutions $\dot{q}$} \label{solutionsdotqC} I now return to the important equations (\ref{eqnsRfinal}). These equations are valid for $q \in D_{q}$, but will later only be considered at a fixed point $q_{0}$, which is identified as the initial value point for the d'Alembertian and variational equations of motion (\ref{Lagnh3}) and (\ref{CC1}), respectively. The initial velocity at $q_{0}$ will be denoted by $\dot{q}_{0}$. For simplicity of notation, I will for the time being consider the equations at an arbitrary fixed point $q \in D_{q}$. Contracting the equations (\ref{eqnsRfinal}) with an arbitrary basis-vector $b_{\nu}(q)$, with $\nu$ in the range $2p+1, \ldots, N$, one obtains, \begin{equation} \label{bRcond} (b_{\nu}(q), a(q))\,||M(q)\dot{q}|| = 0, \; \nu = 2p+1, \ldots, N. \end{equation} From Eq. (\ref{bRcond}) follows that {\em either}, \begin{equation} \label{Rzero} ||M(q)\dot{q}|| = 0, \end{equation} {\em or}, \begin{equation} \label{Rnotzero} ||M(q)\dot{q}|| \neq 0, \end{equation} in which case the following conditions must also be fulfilled, \begin{equation} \label{altRcond} (b_{\nu}(q), a(q)) = 0, \; \nu = 2p+1, \ldots, N. \end{equation} \subsubsection{The case $||M(q)\,\dot{q}|| = 0$.} Consider first the case (\ref{Rzero}) with $2p < N$. Thus $\dot{q} \in ker(M(q)$, or, equivalently, \begin{equation} \label{solqdothom1} \dot{q}^{A} = \sum_{\nu=2p+1}^{N} \gamma_{\nu}\, b_{\nu}^{~A}(q), \;A = 1, 2, \ldots, N. \end{equation} The parameters $\gamma_{\nu}, \nu = 2p+1, \ldots, N$, in the expressions (\ref{solqdothom}) are restricted by the non-holonomic condition (\ref{1nonhol}). Thus, \begin{equation} \label{condam1} 0 = (a(q), \dot{q}) = \sum_{\nu=2p+1}^{N} \gamma_{\nu}\, (a(q), b_{\nu}(q)). \end{equation} I now consider the results above at the fixed initial value point $q_{0}$. The equation (\ref{solqdothom1}) then reads as follows. \begin{equation} \label{solqdothom} \dot{q_{0}}^{A} = \sum_{\nu=2p+1}^{N} \gamma_{0 \nu} \,b_{\nu}^{~A}(q_{0}), \;A = 1, 2, \ldots, N. \end{equation} The parameters $\gamma_{0 \nu}, \, \nu = 2p+1, \ldots,, N$, are restricted by the condition (\ref{condam1}) evaluated at $q_{0}$, \begin{equation} \label{condam} 0 = (a(q_{0}), \dot{q}_{0}) = \sum_{\nu=2p+1}^{N} \gamma_{0 \nu}\, (a(q_{0}), b_{\nu}(q_{0})). \end{equation} However, the condition (\ref{condam}) is nugatory if the vector $a(q_{0})$ is orthogonal to $\ker(M(q))$. There are thus at most $N-2p$ parameters $\gamma_{0 \nu}$ available for the $N-1$ independent initial values obtainable from the initial velocity components $\dot{q}_{0}^{A}, A = 1, 2, \ldots, N$. This means that the system of equations (\ref{solqdothom}), considered as equations for the determination of the parameters $\gamma_{0 \nu}, \nu = 2p+1, \ldots, N$ is over-determined. These equations can be consistent only if the given initial values $\dot{q}_{0}^{A}, A=1, 2,\ldots, N$ satisfy at least $(N-1) - (N-2p) = 2p-1 \geq 1$ consistency conditions, in addition to the constraint (\ref{initrestr}). This contradicts the requirement that one should be able to impose general initial values $\dot{q}_{0}^{A}, \, A = 1, 2, \ldots, N$, for the velocities $\dot{q}^{A}, A = 1, 2, \ldots, N$, except for the restriction (\ref{initrestr}), which follows from the non-holonomic condition (\ref{1nonhol}) at the initial value point $q_{0}$. It has thus been shown that that the initial values $\dot{q}_{0}^{A}, A = 1, 2, \ldots, N$, which satisfy the condition implied by the non-holonomic constraint (\ref{1nonhol}) evaluated at $q_{0}$, can not be freely chosen, apart from the restriction (\ref{initrestr}), if one demands that the d'Alembertian equations (\ref{Lagnh3}) and the variational equations (\ref{CC1}) should have coincident solutions. There is still the possibility that $ker(M(q))$ contains only the zero vector. This is the case if the matrix $M$ is regular, which is possible only if $N$ is an even integer. Then there is only the trivial solution $\dot{q}_{0} = 0$, which naturally is in conflict with the requirement that it should be possible to choose the initial values $\dot{q}_{0 A}, A = 1, 2, \ldots, N$, freely, except for the condition implied by the non-holonomic constraint (\ref{1nonhol}) at the point $q_{0}$. \subsubsection{The case $||M(q)\, \dot{q}|| \neq 0$.} \label{importanteM} It has been shown above that the case (\ref{Rzero}) necessarily leads to contradictions. It remains to consider the case (\ref{Rnotzero}). Then the conditions (\ref{altRcond}) must also be in force. In view of the conditions (\ref{altRcond}), the vector $a(q)$ must lie in the subspace spanned by the basis-vectors $b_{\mu}(q)$, for $\mu = 1, 2, \ldots, 2p$, \begin{equation} \label{repra} a(q) = \sum_{\mu=1}^{p}\left \{ \omega_{2\mu-1}\,b_{2\mu-1}(q) + \omega_{2\mu}\,b_{2\mu}(q)\right \}, \end{equation} where for later convenience we have separated the sum in the representation (\ref{repra}) into two parts, as sums over even and odd indices, respectively. The coefficients $\omega_{2\mu-1}$ and $\omega_{2\mu}$, respectively, for $\mu = 1, \ldots, p$, in the representation (\ref{repra}), are determined by the vector $a(q)$. I will now consider such solutions $z$, say, of the equations (\ref{eqnsRfinal}) which are in the linear span of the basis-vectors $b_{\mu}(q)$ for $\mu = 1, 2, \ldots, 2p$. Thus, \begin{equation} \label{reprz} z = \sum_{\mu=1}^{p} \left \{ \alpha_{2\mu-1}b_{2\mu-1}(q) + \alpha_{2\mu}b_{2\mu}(q) \right \}. \end{equation} Inserting $\dot{q} = z$ in Eqns. (\ref{eqnsRfinal}) one obtains the following equations for the components $z^{A}, A = 1, 2, \ldots, N$, \begin{equation} \label{eqnsMz} \sum_{B=1}^{N} M_{AB}(q)z^{B} = \frac{||M(q)\,z||}{||a(q)||}\,a_{A}(q), \;A = 1, 2, \ldots, N. \end{equation} Contracting the equations (\ref{eqnsMz}) with the basis-vector components $b_{2\nu}^{~~A}(q)$ and using the equations (\ref{Meqeven}), one obtains the following result, \begin{equation} \label{oddcompz} \kappa_{\nu}(q) (b_{2\nu-1}(q), z) = \frac{||M(q)\,z||}{||a(q)||}\, (b_{2\nu}(q), a(q)), \;\nu = 1, \ldots, p. \end{equation} Likewise, contracting the (\ref{eqnsMz}) with the basis vector components $b_{2\nu-1}^{~~~~~A}(q)$, and using the equations (\ref{Meqodd}), one finds a similar expression, \begin{equation} \label{evencompz} \kappa_{\nu}(q) (b_{2\nu}(q), z) = - \frac{||M(q)\,z||}{||a(q)||}\, (b_{2\nu-1}(q), a(q)), \; \nu = 1, \ldots, p. \end{equation} The equations (\ref{oddcompz}) and (\ref{evencompz}) mean that the equations (\ref{eqnsRfinal}) for the vector $z$ have been solved in the sense that the equations (\ref{oddcompz}) and (\ref{evencompz}) express all the components $\alpha_{2\nu-1}$ and $\alpha_{2\nu}$, respectively, in the representation (\ref{reprz}) for $z$, in terms of one unknown scalar quantity, namely $||M(q)\,z||$, and the known components $\omega_{2\nu-1}$ and $\omega_{2\nu}$, respectively, in the representation (\ref{repra}) for the vector $a(q)$. The result obtained above can be summarised as follows, \begin{equation} \label{zfinale1} z^{A}= \frac{||M(q)\, z||}{||a(q)||} \sum_{\nu=1}^{p} \kappa_{\nu}^{-1}(q) \left \{(b_{2\nu}(q), a(q))\,b_{2\nu-1}^{~~~~~ A}(q) - (b_{2\nu-1}(q), a(q))\,b_{2\nu}^{~~ A}(q) \right \}. \end{equation} It should noted that the vector $z$ defined by Eqns. (\ref{zfinale1}) automatically fulfils the following orthogonality condition, \begin{equation} \label{z0ortoa1} (z, a(q)) = 0. \end{equation} So far only the solutions $z^{A}, A = 1, 2, \ldots, N$, of the equations (\ref{eqnsRfinal}) in the linear span of the basis vectors $b_{\mu}(q), \mu = 1, \ldots, 2p$, have been considered. The general solution $\dot{q}$ of Eqns. (\ref{eqnsRfinal}) is a sum of the solution $z$ given above in Eqns. (\ref{zfinale1}), and a general solution $h$, say, of the corresponding homogeneous equations. Thus, \begin{equation} \label{H0} h^{A} := \sum_ {\nu=2p+1}^{N} \gamma_{\nu}\, b_{\nu}^{~A}(q), \end{equation} where the parameters $\gamma_{\nu}, \nu = 2p+1, \ldots, N$ are free parameters. The general solutions of Eqns. (\ref{eqnsRfinal}) for the components $\dot{q}^{A}$ are the following, \begin{equation} \label{qsol1} \dot{q}^{A} := z^{A} + h^{A}, \; A = 1, 2, \ldots, N \end{equation} It should be noted that \begin{equation} \label{Cmnorm} ||M(q)\dot{q}|| = ||M(q)(z+ h)|| = ||M(q)\,z||. \end{equation} One thus finally obtains the following expressions for the general solutions of Eqns. (\ref{eqnsRfinal}), \begin{eqnarray} \label{qdotfinale1} \dot{q}^{A} & = & \frac{||M(q)\,\dot{q}||}{||a(q)||} \sum_{\nu=1}^{p} \kappa_{\nu}^{-1}(q) \left \{(b_{2\nu}(q), a(q))\,b_{2\nu-1}^{~~~~~ A}(q) - (b_{2\nu-1}(q), a(q))\,b_{2\nu}^{~~ A}(q) \right \} \nonumber \\ & + & \sum_ {\nu=2p+1}^{N} \gamma_{\nu}\, b_{\nu}^{~A}(q), \;A = 1, 2, \ldots, N. \end{eqnarray} In view of the conditions (\ref{altRcond}) and (\ref{z0ortoa1}), one finds that the final expressions (\ref{qdotfinale1}) for the components $\dot{q}_{0}^{A}, \, A = 1, 2, \ldots, N$, automatically satisfy the non-holonomic condition (\ref{1nonhol}). I now apply the results above for the initial value $\dot{q}_{0}$ at $q_{0}$. It follows from Eqns. (\ref{qdotfinale1}) that \begin{eqnarray} \label{qdotfinale} \dot{q}_{0}^{A} & = & \frac{||M(q_{0})\,\dot{q}_{0}||}{||a(q_{0})||} \sum_{\nu=1}^{p} \kappa_{\nu}^{-1}(q_{0}) \left \{(b_{2\nu}(q_{0}), a(q_{0}))\,b_{2\nu-1}^{~~~~~ A}(q_{0}) - (b_{2\nu-1}(q_{0}), a(q_{0}))\,b_{2\nu}^{~~ A}(q_{0}) \right \} \nonumber \\ & + & \sum_ {\nu=2p+1}^{N} \gamma_{0 \nu}\, b_{\nu}^{~A}(q_{0}), \;A = 1, 2, \ldots, N. \end{eqnarray} When the initial velocity components $\dot{q}_{0}^{A}, A = 1, 2, \ldots, N$ are given, the only unknown quantities in the equations (\ref{qdotfinale}) are the $N-2p$ parameters $\gamma_{0 \nu}, \nu = 2p+1, \ldots, N$. When $2 \leq 2p < N$, there are altogether $N-1$ independent equations among the $N$ equations in (\ref{qdotfinale}), to be solved for the $N-2p$ parameters $\gamma_{0 \nu}, \nu = 2p+1, \ldots, N$, when the components $\dot{q}_{0}^{A}, A = 1, 2, \ldots, N$ are given. This is an over-determined system of equations for the parameters $\gamma_{0 \nu}, \nu = 2p+1, \ldots, N$, which implies that there must be at least $(N-1) - (N-2p) = 2p-1 \geq 1$ consistency conditions among the components $\dot{q}_{0}^{A}, A = 1, 2, \ldots, N$, in addition to the constraint (\ref{initrestr}). This contradicts the requirement that it should be possible to choose the initial velocity $\dot{q}_{0}$ freely, except for the constraint (\ref{initrestr}). For even $N$ one can have $2p = N$. Then the solutions for $\dot{q}_{0}^{A}, \, A = 1,2,\ldots, N$ are as follows, \begin{equation} \label{qdotfinale2} \dot{q}_{0}^{A} = \frac{||M(q_{0})\,\dot{q}_{0}||}{||a(q_{0})||} \sum_{\nu=1}^{p} \kappa_{\nu}^{-1}(q_{0}) \left \{(b_{2\nu}(q_{0}), a(q_{0}))\,b_{2\nu-1}^{~~~~~ A}(q_{0}) - (b_{2\nu-1}(q_{0}), a(q_{0}))\,b_{2\nu}^{~~ A}(q_{0}) \right \} \end{equation} The $N$ components $\dot{q}_{0}^{A}, \; A = 1, 2, \ldots, N$ in the equations (\ref{qdotfinale2}), are given in terms of one undetermined scalar quantity $||M(q_{0})\,\dot{q}_{0}||$, in addition to the other known quantities in (\ref{qdotfinale2}). This means that there are $N-1$ constraints among the components $\dot{q}_{0}^{A}, \; A = 1, 2, \ldots, N$, which clearly contradicts the requirement that it should be possible to choose the initial velocity $\dot{q}$ freely, except for the constraint (\ref{initrestr}). In the case $||M(q_{0})\,\dot{q}_{0}|| \neq 0$, it has thus been shown that there are at least $2p-1$ consistency conditions on the velocity components $q_{0}^{A}$. There are also $N-2p$ conditions of the form (\ref{altRcond}) evaluated at $q_{0}$, namely the following conditions, \begin{equation} \label{altRcond2} (b_{\nu}(q_{0}), a(q_{0})) = 0, \; \nu = 2p+1, \ldots, N, \end{equation} Taken together there are thus altogether at least $(2p-1) + (N-2p) = N-1$ conditions on the initial values $(q_{0}^{A}, \dot{q}_{0}^{A}), A = 1, 2, \ldots, N$. This is a set of stronger conditions than the consistency conditions involving only the velocity components $\dot{q}_{0}^{A}, A = 1, 2, \ldots, N$. It has thus been shown also in the case $||M(q)_{0})\dot{q}_{0}|| \neq 0$, that that the initial values $\dot{q}_{0}^{A}, A = 1, 2, \ldots, N$, which satisfy the condition (\ref{initrestr}) implied by the non-holonomic constraint (\ref{1nonhol}) evaluated at $q_{0}$, have to satisfy additional consistency conditions, if one demands that the d'Alembertian equations (\ref{Lagnh3}) and the variational equations (\ref{CC1}) should have coincident solutions. \subsection{Conclusion of the proof of the incompatibility theorem} The starting point in of the beginning of the proof in Subsection \ref{preliminary}, is the following set of equations, \begin{equation} \label{eqnsM2} \sum_{B=1}^{N} M_{AB}(q)\,\dot{q}^{B} = \Gamma\,a_{A}(q), \; A = 1, \ldots, N, \end{equation} where $\Gamma$ is a scalar parameter. The equations (\ref{eqnsM2}) are necessary consequences of the assumption that the variational equations (\ref{CC1}) and the d'Alembertian equations (\ref{Lagnh3}) have coincident solutions. It was shown in Subsection \ref{dimension3} in the case $N=3$, that the equations (\ref{eqnsM2}) do not have solutions $\dot{q}^{A}, \, A = 1,2,3$, satisfying the required constraint (\ref{1nonhol}), if this constraint is genuinely non-holonomic, {\em i.e.}, neither integrable as such, nor integrable by means of an integrating factor. Thus the variational equations (\ref{CC1}) and the d'Alembertian equations (\ref{Lagnh3}) are not compatible when $N=3$. When $N\geq4$, the equations (\ref{eqnsM2}) were transformed into the following form in Subsection \ref{dimens4}, \begin{equation} \label{eqnsRfinalC2} \sum_{B=1}^{N} M_{AB}(q)\dot{q}^{B} = \frac{||M(q)\dot{q}||}{||a(q)||}\, a_{A}(q), A = 1, 2, \ldots, N, \end{equation} where the norms $||\ldots||$ occurring above are defined by an appropriate metric tensor $g$. For $N\geq4$ the variational equations of motion (\ref{CC1}), and the d'Alembertian equations of motion (\ref{Lagnh3}), respectively, have been considered to be initial value problems. The implications of Eqns. (\ref{eqnsRfinalC2}) for the initial values $q_{0}$ and $\dot{q}_{0}$ were analysed in Subsection \ref{solutionsdotqC}. It was shown that the equations (\ref{eqnsRfinalC2}) imply that the initial values $\dot{q}_{0}^{A}, \, A = 1,2,\ldots, N$, for the solutions of the variational equations (\ref{CC1}) and d'Alembertian equations (\ref{Lagnh3}), which satisfy the required constraint (\ref{1nonhol}), will have to satisfy additional consistency conditions if the constraint (\ref{1nonhol}) is genuinely non-holonomic. Thus in the cases $N\geq4$, the d'Alembertian and variational equations of motion are incompatible in the sense that in the case of a genuinely non-holonomic constraint, they can not have coinciding solutions with initial velocities $\dot{q}_{0}^{A}, \, A = 1, 2, \ldots, N$, which satisfy the required constraint (\ref{1nonhol}) at $q_{0}$, unless these initial velocities satisfy additional consistency conditions. This concludes the proof of the incompatibility theorem stated in Subsection \ref{incomp}. \section{Discussion} In this paper I have refined and completed an earlier proof \cite{CCTRJMP} of the incompatibility of the variational equations of motion (\ref{CC1}) and the d'Alembertian equations of motion (\ref{Lagnh3}). It was observed in Ref. \cite{CCMN70}, that the proof in Ref. \cite{CCTRJMP} is not valid in the cases $p = 1$ and $N\geq4$, since the conditions of consistency derived in the proof in Ref. \cite{CCTRJMP} are nugatory when $p=1$ and $N\geq4$. This circumstance has now been taken into account in the present proof, in which more precise consistency conditions have been derived than in the earlier proof. In the present proof the formulation respects explicitly the desirable circumstance that the geometry of configuration space need not be Euclidean, but rather {\em e.g.} Riemannian. The tensor machinery necessary in this case, which was developed in Subsection \ref{dimens4}, and in particular in Subsection \ref{digrM}, may seem a little heavy. It is a price one has to pay for the desire to handle a Riemannian configuration space using a classical tensor formulation. The proof in Ref. \cite{CCTRJMP} avoided using a full-fledged tensor formalism, by using essentially only local analysis related to the fixed initial point $q_{0}$ in configuration space for the cases $N\geq4$, and by implicitly using so-called normal- or geodesic co-ordinates \cite {Eisenh} at that point. This means that one was permitted to use a Cartesian metric at the fixed point $q_{0}$ in Ref. \cite{CCTRJMP}. The incompatibility theorem proved in this paper does not mean that one is always unable to associate some variational principle with systems with non-holonomic constraints. It has only been proved that the particular variational equations (\ref{CC1}) are incompatible with the d'Alembertian equations of motion when $N\geq4$, in the sense that these equations do not have solutions with general initial velocities, if the homogeneous constraint used in this paper is genuinely non-holonomic. For $N=3$ a stronger result was proved, namely that the equations of motion in question do not have coinciding solutions at all, if the constraint is genuinely non-holonomic. Conversely, it is by no means guaranteed that the variational equations (\ref{CC1}), and the d'Alembertian equations (\ref{Lagnh3}) have coinciding solutions, if the initial values $\dot{q}_{0}^{A}, \, A = 1, 2, \ldots, N$, satisfy those partly explicit and partly implicit consistency conditions which have been derived in this paper. For instance, the consistency consistency conditions involving only the initial velocities do by no means exhaust all the implications of the equations (\ref{eqnsRfinalC2}) for the initial values $q_{0}$ and $\dot{q}_{0}$, not to mention additional consistency requirements which come into play when one takes into account the full information residing in the equations of motion. An example of reasonably simple yet intricate additional consistency conditions on the initial values $q_{0}$ was given for an important class of solutions of the equations (\ref{eqnsRfinalC2}) in Subsection \ref{importanteM}. \vspace*{2cm} \noindent \large {\bf Acknowledgements} \normalsize I am indebted to the organisers of the "4th International Young Researchers Workshop on Geometry, Mechanics and Control", which took place at the University of Ghent in January 11-13, 2010, and in particular to Dr. Tom Mestdag, for inviting me to this inspiring meeting. I would also like to thank my friend Professor Jerry Segercrantz for an illuminating discussion when I was finalising the manuscript of this paper. \newpage
\section{Introduction} If a liquid is cooled below its melting temperature the liquid will exist in a metastable state until a nucleation event occurs. If the source of nucleation in the undercooled melt is only due to fluctuation phenomena the nucleation is called homogeneous. In classical nucleation theory a spherical shape for the critical nuclei is assumed and its size is determined as a result of competition between the bulk free energy reduction and interfacial energy increase. If $V$ is the volume, $A$ the surface area, $\Delta g$ the chemical free energy change per unit volume and $\gamma$ the specific interfacial energy, the free energy change according to the formation of a new phase is given by $\Delta G = V \Delta g + A \gamma$. For a spherical shape with radius $r$ we thus obtain $\Delta G = 4/3 \pi r^3 \Delta g + 4 \pi r^2 \gamma$. The radius $r^*$ of the critical nucleus must then be such that $r^* = - 2 \gamma / \Delta g$ with the critical free energy of formation of a critical nucleus $\Delta G^* = 16 \pi \gamma^3 / (3 (\Delta g)^2)$. This classical theory has been utilized to interprete kinetics of many phase transformations and have had some success for providing good descriptions on the nucleation kinetics for various systems. On the other hand, nucleation is generally significantly more complicated. The shape might not be spherical due to an anisotropy of the interfacial energy between a nucleus and the bulk phase, which results from the crystallographic nature of a solid nuclei. Furthermore, the bulk properties of small nuclei may differ from bulk values typically obtained from larger samples. To account for these phenomena various new attempts in the context of diffuse interface models have been made to describe nucleation. Such a non-classical theory was pioneered by Cahn and Hilliard \cite{CahnHilliard_JCP_1959}. For subsequent studies, generalizations and specific applications to nucleation, we refer to \cite{GranasyBoerzoenyyiPusztai_PRL_2002,GranasyPusztaiSaylorWarren_PRL_2007,ZhangChenDu_PRL_2007,WarrenPusztaiKoernyeiGranasy_PRB_2009} and the references therein. In these studies an order parameter is used to distinguish between the nucleus and the bulk phase. Since nucleation takes place by overcoming the minimum energy barrier, a critical nucleus is defined as the order parameter fluctuation which has the minimum free energy increase among all fluctuations which lead to nucleation. Therefore, the critical nucleus can be found by computing the saddle points of the energy functional of the order parameter, that has the highest energy in the minimum energy path (MEP), which is the path whose highest energy is the lowest among all possible paths. This is consistent with the large deviation theory which states that the most probable path passes through the saddle point in the large time limit. An efficient numerical approach for finding minimum energy paths and saddle points, the so-called string methods (SM), has been introduced in \cite{Eetal_PRB_2002}. The method is related to the nudged elastic band (NEB) method \cite{HenkelmanJonsson_JCP_2000}. Other approach are e.g. the minimax method, which as been used in \cite{ZhangChenDu_JSC_2008} or a phase-field type approach, as used in \cite{Iwamatsu_JCP_2009} in the context of nucleation. We will here apply a simplified string method (SSM) \cite{Eetal_JChemP_2007} but not on an underlying diffuse-interface model but a more detailed phase-field-crystal model \cite{Elderetal_PRL_2002}, which accounts for the discrete effects on the small length scales involved in nucleation. The outline of the paper is as follows: In Sec. 2 we introduce the phase-field-crystal model as a local approximation of a classical dynamic density functional theory. In Sec. 3 we describe the used string method. In Sec. 4 we discuss implementational issues. In Sec. 5 we show results for homogeneous nucleation. Conclusions are drawn in Sec. 6. \section{Phase-field-crystal model} The phase-field-crystal (PFC) model is by now widely used in order to describe solid-state phenomena on atomic length scales. The PFC model was first developed in \cite{Elderetal_PRL_2002} and then subsequently applied to many situations like interfaces \cite{Athreyaetal_PRE_2007,BackofenVoigt_JPCM_2009}, polycrystalline pattern formation \cite{Wuetal_PRB_2007, Goldenfeldetal_PRE_2005}, commensurate-incommensurate transitions \cite{Achimetal_PRE_2006}, edge dislocations \cite{Berryetal_PRE_2006}, grain boundary pre-melting \cite{Plappetal_PRB_2008}, colloidal solidification \cite{vanTeeffelenetal_PRE_2009} and dislocation dynamics \cite{BackofenBernalVoigt_IJMR_2010}. The model resolves the atomic-scale density wave structure of a polycrystalline material and describes the defect-mediated evolution of this structure on time scales orders of magnitude larger than molecular dynamic (MD) simulations. In its simplest form the PFC model results from the energy \begin{equation} \label{eq:0} F[\varphi] = \int_\Omega - |\nabla \varphi|^2 + \frac{1}{2} (\Delta \varphi)^2 + f(\varphi) \; dx \end{equation} with $f(\varphi) = \frac{1}{2}(1 - \epsilon)\varphi^2 + \frac{1}{4} \varphi^4$ a potential, $\varphi$ the number density and $\epsilon$ a parameter determining the approximation of the liquid structure factor ~\cite{Elderetal_PRL_2002}. Comparing the energy with a classical phase-field type energy, e.g. $\int_\Omega \frac{\delta}{2}|\nabla \phi|^2 + \frac{1}{\delta} g(\phi) \; dx$ for an order parameter $\phi$, with $\delta$ a length scale determining the width of a diffuse interface and $g(\phi)$ a double well potential, the difference is in the sign of the gradient term and the additional higher order term. The negative sign in the gradient term favors a changes in $\varphi$, whereas the higher order term favors to suppress such changes. The competition between both terms introduces a fixed length scale for which the energy will be minimized. This length scale is used to model the periodicity of a crystal lattice. The formulation used here favors a hexagonal closed packed structure in two dimensions. Due to the underlying periodicity, several solid state phenomena such as elasticity, plasticity, anisotropy and multiple grain orientations are naturally present in the formulation. The dynamic law constructed to minimize the free energy follows as the $H^{-1}$-gradient flow of the energy \begin{equation} \label{eq:1} \partial_t \varphi = \Delta \mu \end{equation} with chemical potential $\mu = \frac{\delta F[\varphi]}{\delta \varphi}$ and the variational derivative given by \[ \frac{\delta F[\varphi]}{\delta \varphi} = \Delta^2 \varphi + 2 \Delta \varphi + f^\prime(\varphi). \] This defines the PFC model and its evolution is by construction towards a (meta) stable state. Although this formulation is phenomenologically, the PFC model can be derived starting from a Smoluchowski equation via dynamic density functional theory using various approximations \cite{Elderetal_PRE_2007,vanTeeffelenetal_PRE_2009}. With an appropriate parameterization it thus provides also a quantitative atomic theory, operating on diffusive time scales. Within new developments \cite{vanTeeffelenetal_PRE_2009,Jaatinenetal_PRE_2009} quantitative agreement in computed properties could be achieved using the PFC model for various materials. It thus provides an ideal model to study nucleation. \section{Minimum energy path} \subsection{Definition} The dynamics shown in eq. (\ref{eq:1}) describe the evolution towards equilibrium of a single state $\varphi$ in phase space according to the generalized thermodynamic force $\Delta \mu$. But in order to characterize nucleation the most likely transition path between (meta) stable states has to be identified. In the description of such a non equilibrium process, the minimum energy path (MEP) plays a crucial role. The MEP is a path in phase space that connects (meta) stable states. A path in phase space is thereby defined as \begin{eqnarray*} \gamma_c = \left\{ \varphi_\alpha : \alpha \in [0,1] \right\} \end{eqnarray*} with $\alpha$ a parameterization of the path. For the MEP the generalized thermodynamic force $\Delta \mu$ is tangential to this path: \begin{eqnarray} \left( \Delta \mu \right)^{\perp} =0 \end{eqnarray} Thus, using the dynamics in eq. (\ref{eq:1}), any state of the MEP will always evolve along the MEP towards a stable state. That is, the MEP is a real reaction path in phase space and along the MEP the energy is defined by eq. (\ref{eq:0}). The local energetic maxima and minima along the MEP can be used to determine the nucleation barrier and the critical nucleus. The string method (SM) is been designed to find the MEP. It evolves a given chain of states towards the MEP, see Fig. \ref{fig:1} for illustration. \begin{figure}[htb] \noindent \center \includegraphics*[angle = -0, width = 0.5 \textwidth ]{figures/ref_iso_allInOne_nice} \begin{center} \begin{minipage}{0.9\textwidth} \caption[short figure description]{ Schematic sketch of a free energy surface of a two dimensional phase space. Two (meta) stable states (valleys) are separated by a saddle point. The single states (circles) are evolved by thermodynamic forces (arrows). The initial string (straight line) is evolved by the string method towards the MEP. \label{fig:1} } \end{minipage} \end{center} \end{figure} \subsection{String method} A path in phase space $\gamma_c$ is represented by a discrete set of states $\varphi_i$, denoted by \begin{eqnarray*} \gamma = \left\{ \varphi_i : i = 0,1, \ldots, N \right\} = \left\{ \varphi_i \right\}. \end{eqnarray*} The length of the path is defined by \begin{eqnarray*} L(\gamma)=\sum_{i=0}^{N-1} \left| \varphi_{i+1}-\varphi_i \right| \quad {\rm and } \quad \left| \varphi \right| = \int \left| \varphi(r) \right| \,dr \end{eqnarray*} where $|\cdot|$ measures the distance between two states and is defined by the $L^2$-norm. Thus, the normalized tangent $\hat{t}$ to the path at state $\varphi_i$ may be calculated as $\hat{t_i}=\frac{\varphi_{i+1}-\varphi_i}{\left|\varphi_{i+1}-\varphi_i\right|}$. The idea behind the SM is to minimize the free energy of all single states according to \begin{eqnarray} \varphi_i^{n+1}=\tau \Delta \mu + \varphi_i^n. \end{eqnarray} but restricting the evolution to orthonormal direction of the path. In addition a tangential force is included in order to keep the quality of the representation of the path by the string. \begin{eqnarray} \label{eq-SM} \gamma^n \rightarrow& \gamma^{n+1}&= \left\{ \varphi_i^{n+1} \right\} \\ \nonumber {\rm with}&\varphi_i^{n+1} &=\tau \left( \Delta \mu \right)^\perp+ \varphi_i^n +\lambda_i \hat{t} \end{eqnarray} The Langrang multipliers $\lambda_i$ are e.g. uniquely determined by forcing an equidistant distribution of states along the path. $\tau$ is a fictitious timestep and controls the velocity of evolution. Eq. (\ref{eq-SM}) defines the String Method. It is easily seen that the MEP $\gamma_{\rm MEP}$ is an invariant according to the dynamics of SM. By definition of the MEP, the thermodynamic force is only tangential to the path. The constraints introduced by the tangential force do not alter the path, but only reparameterize the representation of $\gamma$. Thus, $\gamma_{\rm MEP}^n$ and $\gamma_{\rm MEP}^{n+1}$ represents the same path in phase space. In order to implement the SM the thermodynamic force has to be calculated and projected to the orthogonal direction of the path. Furthermore the Langrange multipliers have to be calculated. In order to simplify the calculation the SM can be divided in two steps leading to a Simplified String Method (SSM). First the string is evolved due to the thermodynamic force and then the path is reparameterized, see \cite{Eetal_JChemP_2007}. That is, the states representing the path $\left\{\tilde{\varphi}_i\right\}$ are replaced by equally distant states, that represent the same path $\left\{\varphi_i\right\}$. This new states are constructed by interpolation between the original states $\left\{\tilde{\varphi_i}\right\}$. As in every evolution step there is a parameterization step, it is no longer necessary to project the thermodynamic force. Thus the SSM is defined by two steps: \begin{enumerate} \item Evolution step: \begin{eqnarray} \gamma^n \rightarrow& \tilde{\gamma}^{n+1}&= \left\{ \tilde{\varphi}_i^{n+1} \right\} \\ \nonumber {\rm with}&\tilde{\varphi}_i^{n+1} &= \tau \Delta \mu + \varphi_i^n \end{eqnarray} \item Reparameterization step: \begin{eqnarray} &\tilde{\gamma}^{n+1} \rightarrow \gamma^{n+1}=\left\{\varphi_i^{n+1}\right\}\\ \nonumber {\rm such~that}& \left| \varphi_{i+1}^{n+1}-\varphi_i^{n+1} \right| = \frac{L(\tilde{\gamma}^{n+1})}{N-1} \quad i = 0,1,2, \ldots,N-1 \end{eqnarray} and $\gamma^{n+1}$ and $\tilde{\gamma}^{n+1}$ representing the same path in phase space. \end{enumerate} Here the reparameterization is done to force equidistant states on the path. However, the reparameterization may also be changed to account for problem specific details, e.g. to get finer representation at the saddle point. The advantage of SSM over SM is that the thermodynamic force has not to be projected. Additionally it is shown that this modification leads to a more stable and accurate method \cite{Eetal_JChemP_2007}. \subsection{Fixed length Simplified String Method} In the above defined SSM an initial path in phase space is evolving towards the MEP. The first and the last state representing the path thereby converge to different (meta) stable states. The saddle point or here the critical nucleus is defined by the state of highest energy in the MEP. If there is only one energetic maximum and the saddle point is well defined, the MEP could be calculated easily by just solving the time evolution of a small perturbation of the critical nucleus towards the stable states according to eq. (\ref{eq:1}). Thus, only two time dependent simulations have to be done. Therefore, it is enough to find the saddle point to reconstruct the whole MEP efficiently. In order to concentrate the simulation effort to find only the saddle point, we introduce a Fixed Length Simplified String Method (FLSSM). The total length of the string is restricted by the reparameterization step. Assume that we allow a maximal length of the string, $L_{\rm fixed}$. For simplicity we also assume that the first state converges towards a stable state. Then, the reparameterization can always project the states back to a string beginning with the first state with length $L_{\rm fixed}$. The last state is not converging to a meta stable state, but might be some unstable state but within a different basin. This state can be used to reconstruct the whole MEP by a time evolution according to eq. (\ref{eq:1}). The same idea can be used at the same time on both sides of the saddle point, by fixing a state near the saddle point and restricting the length on both sides of the path. As we need some information about the string length and the position of the saddle point, we use the FLSSM in order to refine and improve accuracy of a MEP which was calculated by the standard SSM with only a few states. The method can be viewed as an adaptive approach which efficiently finds the saddle point within a given tolerance. \subsection{Implementation of a Fixed Length Simplified String Method for PFC} The string is a set of density distribution $\left\{\varphi_i\right\}$. As we consider a closed system and have mass conserving dynamics, we have to restrict the possible states representing a string to the same mean density, $\bar{\varphi} = \int \varphi_i(r) \, dr$ for all $i$. For every state the standard dynamics has to be solved according to eq. (\ref{eq:1}). The partial differential equation of 6th order is splitted into a set of three second order equations: \begin{eqnarray*} \partial_{t} \varphi_i &=& \Delta \mu \\ \mu &=& \Delta v + 2 \Delta \varphi_i + f^\prime(\varphi_i) \\ v &=& \Delta \varphi_i \end{eqnarray*} for which a stable semi-implicit finite element discretization is introduced in \cite{Backofenetal_PM_2007}. We use this approach but with higher order elements. The algorithm is implemented in the adaptive finite element toolbox AMDiS \cite{Veyetal_CVS_2007}. The fictitious timestep is adjusted such that the reparameterization step can be done mostly considering only neighbouring states and such that the evolved state is substantial different from the previous one. In this work we use linear interpolation between the states. We define the length of the string up to state M in analogy to $L(\gamma)$ for the whole string by $L_M(\gamma)= \sum_{i=0}^{M-1} \left|\varphi_{i+1}-\varphi_i \right|$. The distance between states after reparameterization is $\bar{l}=\frac{L(\tilde{\gamma})}{N-1}$. Then the reparameterized state $\varphi_i$ at $L^*=i\bar{l}$ is constructed by linear interpolation using the neighboring states of $\varphi_i$ from $\tilde{\gamma}$. \begin{eqnarray} \varphi_i= \tilde{\varphi}_k+(\tilde{\varphi}_{k+1}-\tilde{\varphi}_{k})\alpha \\ \alpha = \frac{L^*-L_k(\tilde{\gamma})}{\left|\tilde{\varphi}_{k+1}-\tilde{\varphi}_k \right|} \quad {\rm and} \quad L_k(\tilde{\gamma}) \leq L^* < L_{k+1}(\tilde{\gamma}) \end{eqnarray} The FLSSM is implemented in a parallel way. That is, every state define a process and the evolution step is calculated in parallel. The result is then send to the nodes of the neighbouring states. In order to avoid complicated communication between the processes, the reparameterization step is further simplified. The linear interpolation described above is used if the reparameterized state is in between the neighbouring states. If not, the reparameterized state is set to one of the neighbouring sites. \begin{eqnarray} \varphi_i= \left\{ \begin{array}{ll} \varphi_{i-1},& L^* < L_{i-1}(\tilde{\gamma}) \\ {\rm linear~interpolation},& L_{i-1}(\tilde{\gamma}) \leq L^* \leq L_{i+1}(\tilde{\gamma}) \\ \varphi_{i+1},& L^* > L_{i+1}(\tilde{\gamma}) \end{array} \right. \end{eqnarray} This does not alter the MEP but only the dynamic of the string in phase space, as long as we choose the fictitious timestep in a way that for pathes near the MEP at the end of simulation always linear interpolation between neighbouring sites can be used. We define convergence of the method if the string changes only in tangential direction. This is ensured by two criteria. First, the change in string length and the change in energy in every state after reparameterization is below a given tolerance which ensures no change of the string in normal direction. The second criteria accounts for the change in tangential direction and ensures the evolution towards the (meta) stable states. Therefore the fictitious time step is adjusted to evolves the states substantially far but still allowing for linear interpolation with $0.3 < \alpha < 0.7$. If the second criteria is not fullfilled, the fictitious time step is adapted. Due to small thermodynamic forces near the saddle point, the second criteria may be relaxed for states very close to the saddle point. \section{Results} We consider as a proof of concept the nucleation of a crystal grain in an undercooled liquid. The parameters needed in the PFC model, eq. (\ref{eq:1}), are the mean value of the density, $\bar{\varphi}=\int \varphi(r) \, dr$ and the parameter $\epsilon$, which can be interpreted as a driving force of the phase change, e.g. undercooling \cite{Majaniemietal_PRE_2009,Yuetal_2010} or strength of interaction \cite{vanTeeffelenetal_PRE_2009}. In our example we choose parameters in the coexistence region of the phase diagram $(\epsilon, \bar{\varphi}) = (-0.289, -0.345)$. For this parameter the liquid is a meta stable state. The stable state is a grain in coexistence with liquid. The grain is slightly anisotropic and there is a small density difference between crystal and liquid, see \cite{BackofenVoigt_JPCM_2009}. In order to calculated the MEP an initial string $\gamma^0=\left\{\varphi_i^0\right\}$ has to be defined such that the mean density of every state is equal $\bar{\varphi}=\bar{\varphi}_i^0$ and that the first and the last state evolve towards two different (meta) stable states. In our work, we use two different initial strings to demonstrate that the obtained MEP is independent of the initial configuration. In the first example we set the first state to liquid $\varphi_0(x)=\bar{\varphi}$ and the last state to the equilibrium shape of the grain. The states in between are then constructed by linear interpolation. In the second example, every state was set homogeneous and disturbed by white noise $\eta$ such that $\varphi_i(x)=\bar{\varphi}+S_i \eta$. The strength of the noise $S_i$, was linearly increased starting from $0$ to represent the liquid state towards a large value, which ensures evolution within the coexistence regime. Both initial strings converge to the same MEP shown in Fig.~\ref{fig-MEP}. In order to proof stability of the obtained MEP every state was disturbed independently by some random field and than taken as a initial string to recalculate the MEP. \begin{figure}[htb] \noindent \center \includegraphics*[angle = -0, width = 0.7 \textwidth ]{figures/rectpath_00_nice2} \begin{center} \begin{minipage}{0.9\textwidth} \caption[short figure description]{ Free energy along the MEP. The free energy is plotted relative to the liquid state at normalized string length, l=0. The path is discretized by 94 states. At l=1 the stable state is reached. \label{fig-MEP} } \end{minipage} \end{center} \end{figure} The first state corresponds to liquid and the last to a grain in coexistence with the liquid. The grain equilibrium is energetically favourable compared to the liquid and is the stable state in phase space. The liquid state is meta stable. The nucleation barrier or the saddle point is found at normalized string length of approx. $0.04$. States right to the saddle point correspond to growing crystallites and left to melting crystallites. The string was discretized by 94 states which are equally distributed, so the region around the nucleation barrier is resolved only by 10 states. In order to get a better resolution the FLSSM is used. The length of the string is therefore restricted to $\frac{1}{6}$ of the original length of the MEP and is rediscretized by 46 states, which are constructed by linear interpolation of the calculated MEP. We can view this as an adaptive method to increase the accuracy of the calculated saddle point or a proof that the obtained saddle point is independent of the used parameterization of the string. In our example this independency is shown. Fig. \ref{fig-MEP2} shows the obtained nucleation barrier $\Delta E$ and critical nucleus. \begin{figure}[htb] \noindent \center \includegraphics*[angle = -0, width = 0.7 \textwidth ]{figures/rectpath_len2_00_nice3} \begin{center} \begin{minipage}{0.9\textwidth} \caption[short figure description]{ Detailed MEP around the saddle-point. Closed symbols indicate the MEP calculated by SSM as in Fig. \ref{fig-MEP}. Open symbols show the result achieved by restricting the string length to $\frac{1}{6}$ and using FLSSM. \label{fig-MEP2} } \end{minipage} \end{center} \end{figure} The critical nucleus is defined by the state indicated by (c), $\varphi_{\rm c}$. (b) indicates $\varphi_{\rm b}$ a sub critical state, which most likely will melt. (d) - (f) indicate states $\varphi_{\rm d}$ - $\varphi_{\rm f}$ which correspond to super critical states which will solidify. \begin{figure}[htb] \noindent \center \begin{tabular}{cccc} b & \includegraphics[width = 0.2 \textwidth ]{figures/grain_len_6_s} & c & \includegraphics[width = 0.2 \textwidth ]{figures/grain_len_13_s} \\ d & \includegraphics[width = 0.2 \textwidth ]{figures/grain_len_20_s} & e & \includegraphics[width = 0.2 \textwidth ]{figures/grain_len_40_s}\\ f & \includegraphics[width = 0.2 \textwidth ]{figures/grain_len_63_s} & & \includegraphics[width = 0.2 \textwidth ]{figures/grain_rough_95_s} \end{tabular} \begin{center} \begin{minipage}{0.9\textwidth} \caption[short figure description]{ Various states at the MEP. The labels (b)-(f) indicate the different state $\varphi_{\rm b}$ - $\varphi_{\rm f}$. \label{fig-grain} } \end{minipage} \end{center} \end{figure} In Fig.~\ref{fig-grain} the density field of the labeled states are shown. The critical nucleus is a hexagonal cluster with only seven maxima. A small perturbation of this state will lead either to growth towards the equilibrium shape or to melting. The grow is symmetric and can be seen more quantitative in Fig.~\ref{fig-prof}, which shows the density profile along the x-axis in the various states of the growth process. The density plot shows that the maximum amplitude of the critical nucleus is smaller than in the final bulk state. This can correspond to defects in the crystal, as we consider here only a mean-field description, or weaker ordering of particles. In both cases this shows that the critical nucleus has different structure and bulk energy than the corresponding bulk state. Nucleation thus begins with a disturbance that reflects the crystal structure but has a small amplitude. During growth the spatial size of the initial fluctuation and the amplitude increases. At state $\varphi_{\rm e}$ the maximum amplitude is equal to the bulk value and does not increase anymore. After this state the grain begins to grow only along the solid liquid phase boundary. \begin{figure}[htb] \noindent \center \includegraphics*[angle = 0, width = 0.7 \textwidth ]{figures/cuts_bw} \begin{center} \begin{minipage}{0.9\textwidth} \caption[short figure description]{ Density profile of grains at the MEP. The labels (b)-(f) indicate the different state $\varphi_{\rm a}$ - $\varphi_{\rm f}$, see~Fig.~\ref{fig-MEP2}. \label{fig-prof} } \end{minipage} \end{center} \end{figure} These computations indicate that the subcritical and supercritical grains around the critical nucleus are different to ideal bulk crystals. In addition to the non-spherical shape of the nucleus, which is not considered in classical nucleation theory, such size dependent properties are also not considered in full detail in classical phase field approaches for nucleation. \section{Conclusion} A phase-field-crystal model is used to determine nucleation barriers and the critical nucleus in homogeneous nucleation. The results obtained indicate details of the nucleation process which are not considered in classical nucleation theory but also cannot be addressed in full detail with classical phase field models. The obtained size of the critical nucleus, which here consists only of 7 atoms furthermore asks for an atomistic description. Even if only phenomenological values are used in the computation, the described method gives a proof of concept. The described String Method is independent of the parameterization of the underlying evolution model and thus will allow also to be used for specific materials. With the described implementational details of the String Method, concerning parallel processing and adaptive concepts we believe the approach to be applicable also in three dimensions. \vspace*{0.5cm} {\bf Acknowledgments} We would like to thank E. Vanden-Eijnden for valuable discussions. RB and AV acknowledges support of the DFG via Vo899/7-2 within SPP 1296. \section*{References}
\section{Introduction To briefly set the stage for this paper, let $R>0$, introduce the strip $S_{2 \pi}=\{z\in{\mathbb{C}}\,|\, 0\leq \Re(z) < 2 \pi\}$, and consider the boundary trace map \begin{equation} \label{1.1} \ensuremath{\ga_{\te_0,\te_R}} \colon \begin{cases} C^1({[0,R]}) \rightarrow {\mathbb{C}}^2, \\ u \mapsto \begin{bmatrix} \cos(\theta_0)u(0) + \sin(\theta_0)u'(0)\\ \cos(\theta_R)u(R) - \sin(\theta_R)u'(R) \end{bmatrix}, \end{cases} \quad \theta_0, \theta_R\in S_{2 \pi}, \end{equation} where ``prime'' denotes $d/dx$. In addition, assuming that \begin{equation} V\in L^1((0,R); dx) \label{1.2} \end{equation} (we emphasize that $V$ is not assumed to be real-valued for most of this paper), one can introduce the family of one-dimensional Schr\"odinger operators $\ensuremath{H_{\te_0,\te_R}}$ in $L^2((0,R); dx)$ by \begin{align} & \ensuremath{H_{\te_0,\te_R}} f= -f'' + Vf, \quad \theta_0, \theta_R\in S_{2 \pi}, \notag \\ & f\in\text{\rm{dom}}(\ensuremath{H_{\te_0,\te_R}})=\big\{ g \in L^2((0,R); dx)\,\big|\, g, g'\in AC({[0,R]}); \, \ensuremath{\ga_{\te_0,\te_R}} (g)=0; \label{1.3} \\ & \hspace*{6.75cm} (-g''+Vg)\in L^2((0,R); dx)\big\}, \notag \end{align} were $AC([0,R])$ denotes the set of absolutely continuous functions on $[0,R]$. Assuming that $z\in{\bbC\backslash\sigma(\ensuremath{H_{\te_0,\te_R}})}$ (with $\sigma(T)$ denoting the spectrum of $T$) and $\theta_0, \theta_R \in S_{2 \pi}$, we recall that the boundary value problem given by \begin{align} -&u'' + Vu=zu,\quad u, u'\in AC([0,R]), \label{1.4}\\ &\ensuremath{\ga_{\te_0,\te_R}}(u)=\begin{bmatrix}c_0\\ c_R \end{bmatrix}\in{\mathbb{C}}^2, \label{1.5} \end{align} has a unique solution denoted by $u(z,\cdot)=u(z,\cdot\,;(\theta_0,c_0),(\theta_R,c_R))$ for each $c_0, c_R\in{\mathbb{C}}$. To each boundary value problem \eqref{1.4}, \eqref{1.5}, we now associate a family of \emph{general boundary data maps}, $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z) : {\mathbb{C}}^2 \rightarrow {\mathbb{C}}^2$, for $\theta_0, \theta_R, \theta_0^{\prime},\theta_R^{\prime}\in S_{2 \pi}$, where \begin{align}\label{1.6} \begin{split} \ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z) \begin{bmatrix}c_0\\ c_R \end{bmatrix} &= \ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z) \big(\ensuremath{\ga_{\te_0,\te_R}}(u(z,\cdot\ ;(\theta_0,c_0),(\theta_R,c_R)))\big) \\ &= \ensuremath{\ga_{\te_0^{\prime},\te_R^{\prime}}}(u(z,\cdot\ ;(\theta_0,c_0),(\theta_R,c_R))). \end{split} \end{align} With $u(z,\cdot)=u(z,\cdot\ ;(\theta_0,c_0),(\theta_R,c_R))$, then $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z) $ can be represented as a $2\times 2$ complex matrix, where \begin{align}\label{1.7} \ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z) \begin{bmatrix}c_0\\ c_R \end{bmatrix} &= \ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z) \begin{bmatrix} \cos(\theta_0)u(z,0) + \sin(\theta_0)u'(z,0)\\[1mm] \cos(\theta_R)u(z,R) - \sin(\theta_R)u'(z,R) \end{bmatrix} \notag \\ &= \begin{bmatrix} \cos(\theta_0^{\prime})u(z,0) + \sin(\theta_0^{\prime})u'(z,0)\\[1mm] \cos(\theta_R^{\prime})u(z,R) - \sin(\theta_R^{\prime})u'(z,R) \end{bmatrix}. \end{align} The map $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z)$ represents the principal object studied in this paper. We prove in Section \ref{s2} that $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z)$ is well-defined for $z\in{\mathbb{C}}\backslash\sigma(\ensuremath{H_{\te_0,\te_R}})$, that is, it is invariant with respect to a change of basis of solutions of \eqref{1.4}, derive its basic properties (cf.\ Corollary \ref{c2.4}), and derive the explicit representation \eqref{2.20}, \eqref{2.20a} in terms of a distinguished basis of solutions \eqref{2.5}. In Section \ref{s3}, we relate a special case of $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z)$, given by the generalized Dirichlet-to-Neumann maps $\ensuremath{\La_{\te_0,\te_R}} (z) = \Lateq (z)$, to Weyl--Titchmarsh $m$-functions, derive its asymptotic behavior as $z$ tends to infinity, and most importantly, derive an explicit representation of the boundary data maps $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z)$ in terms of the resolvent of the underlying Schr\"odinger operator $\ensuremath{H_{\te_0,\te_R}}$ and the associated boundary trace maps, \begin{align} \begin{split} \ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z) S_{\theta_0' -\theta_0,\theta_R'-\theta_R} =\gamma_{\theta_0',\theta_R'} \big[\gamma_{\overline{\theta_0'},\overline{\theta_R'}} ((\ensuremath{H_{\te_0,\te_R}})^* - {\overline z} I)^{-1}\big]^*,& \label{1.8} \\[1mm] \theta_0, \theta_R, \theta_0', \theta_R' \in S_{2 \pi}, \; z\in{\mathbb{C}}\backslash\sigma(\ensuremath{H_{\te_0,\te_R}}),& \end{split} \end{align} with $S_{\alpha,\beta}$ denoting the $2 \times 2$ diagonal matrix $S_{\alpha,\beta}= {\rm diag}\big(\sin(\alpha), \sin(\beta)\big)$. Theorem \ref{t4.1}, the principal result in Section \ref{s4}, then centers around the following linear fractional transformation relating the boundary data maps $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z)$ and $\ensuremath{\La_{\de_0,\de_R}^{\de_0^{\prime},\de_R^{\prime}}}(z)$, \begin{align} \begin{split} \ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}}(z) &= \big(S_{\delta_0^\prime-\delta_0,\delta_R^\prime-\delta_R}\big)^{-1} \big[S_{\delta_0^\prime-\theta_0^\prime,\delta_R^\prime-\theta_R^\prime}+ S_{\theta_0^\prime-\delta_0,\theta_R^\prime-\delta_R}\ensuremath{\La_{\de_0,\de_R}^{\de_0^{\prime},\de_R^{\prime}}}(z) \big] \label{1.9} \\ & \quad \times \big[S_{\delta_0^\prime-\theta_0,\delta_R^\prime-\theta_R}+ S_{\theta_0-\delta_0,\theta_R-\delta_R}\ensuremath{\La_{\de_0,\de_R}^{\de_0^{\prime},\de_R^{\prime}}}(z) \big]^{-1}S_{\delta_0^\prime-\delta_0,\delta_R^\prime-\delta_R}, \end{split} \end{align} assuming $\theta_0, \theta_R, \theta_0^{\prime},\theta_R^{\prime}, \delta_0,\delta_R,\delta_0^\prime,\delta_R^\prime\in S_{2 \pi}$, $\delta_0^\prime-\delta_0\ne 0 \, \text{\rm mod} (\pi)$, $\delta_R^\prime-\delta_R \ne 0 \, \text{\rm mod} (\pi)$, and $z\in{\mathbb{C}}\backslash\big(\sigma(H_{\theta_0,\theta_R})\cup \sigma(H_{\delta_0,\delta_R})\big)$. The linear fractional transformation \eqref{1.9} then is a major ingredient in our proof that $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (\cdot) S_{\theta_0'-\theta_0,\theta_R'-\theta_R}$ is a $2 \times 2$ matrix-valued Herglotz function (i.e., analytic on ${\mathbb{C}}_+$, the open complex upper half-plane, with a nonnegative imaginary part) in the special case where $\ensuremath{H_{\te_0,\te_R}}$ is self-adjoint. (In this case, one necessarily assumes that $\theta_0, \theta_R, \theta_0', \theta_R' \in [0,2 \pi)$ and that $V$ is real-valued.) In addition, we derive the Herglotz representation of $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (\cdot) S_{\theta_0'-\theta_0,\theta_R'-\theta_R}$ in terms of a $2 \times 2$ matrix-valued measure on ${\mathbb{R}}$. The principal result proved in Section \ref{s6} then concerns Krein-type resolvent formulas explicitely relating the resolvents of $\ensuremath{H_{\te_0,\te_R}}$ and $H_{\theta_0',\theta_R'}$. A typical result to be proved in Theorrem \ref{t6.3} is of the form \begin{align} & (H_{\theta_0',\theta_R'} -z I)^{-1} = (\ensuremath{H_{\te_0,\te_R}} -z I)^{-1} - \big[\gamma_{\overline{\theta_0^{\prime}},\overline{\theta_R^{\prime}}} ((\ensuremath{H_{\te_0,\te_R}})^* - {\overline z} I)^{-1}\big]^* S_{\theta'_0-\theta'_0,\theta'_R-\theta_R}^{-1} \notag \\ & \hspace*{5.7cm} \times \Big[\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z)\Big]^{-1} \big[\gamma_{\theta_0^{\prime},\theta_R^{\prime}}(\ensuremath{H_{\te_0,\te_R}} - z I)^{-1}\big], \notag \\ & \quad \theta_0, \theta_R, \theta_0', \theta_R' \in S_{2 \pi}, \; \theta_0 \neq \theta_0', \; \theta_R \neq \theta_R', \; z\in{\mathbb{C}}\big\backslash\big(\sigma(\ensuremath{H_{\te_0,\te_R}})\cup \sigma(H_{\theta_0',\theta_R'})\big). \label{1.10} \end{align} Formula \eqref{1.10} demonstrates why $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}}$ is the ideal object for Krein-type resolvent formulas. Finally, in Section \ref{s7}, we describe some additional connections between $\ensuremath{\La_{\te_0,\te_R}} (z)$ and the Green's function $G_{\theta_0,\theta_R} (z,x,x')$ of $\ensuremath{H_{\te_0,\te_R}}$, and then point out some interesting differences compared to the standard $2 \times 2$ Weyl--Titchmarsh matrix associated with $\ensuremath{H_{\te_0,\te_R}}$. For classical as well as recent fundamental literature on Weyl--Titchmarsh operators (i.e., spectral parameter dependent Dirichlet-to-Neumann maps, or more generally, Robin-to-Robin maps, resp., Poincar\'e--Steklov operators), relevant in the context of boundary value spaces (boundary triples, etc.), we refer, for instance, to \cite{ABMN05}, \cite{AB09}, \cite{AP04}, \cite{BL07}, \cite{BMN08}, \cite{BMN00}--\cite{BGP08}, \cite{DHMS00}-- \cite{DM95}, \cite{GKMT01}--\cite{GMZ07}, \cite{GT00}, \cite[Ch.\ 3]{GG91}, \cite{Gr08a}, \cite[Ch.\ 13]{Gr09}, \cite{KO77}--\cite{KS66}, \cite{LT77}, \cite{MM06}, \cite{Ma04}, \cite{Pa87}, \cite{Pa02}, \cite{Po04}, \cite{Po08}, \cite{PR09}, \cite{Ry07}--\cite{Ry10}, \cite{TS77} and the references cited therein. Finally, we briefly summarize some of the notation used in this paper: Let ${\mathcal H}$ be a separable complex Hilbert space, $(\cdot,\cdot)_{{\mathcal H}}$ the scalar product in ${\mathcal H}$ (linear in the second argument), and $I_{{\mathcal H}}$ the identity operator in ${\mathcal H}$. Next, let $T$ be a linear operator mapping (a subspace of) a Banach space into another, with $\text{\rm{dom}}(T)$ and $\ker(T)$ denoting the domain and kernel (i.e., null space) of $T$. The spectrum of a closed linear operator in ${\mathcal H}$ will be denoted by $\sigma(\cdot)$. The Banach space of bounded linear operators on ${\mathcal H}$ is denoted by ${\mathcal B}({\mathcal H})$, the analogous notation ${\mathcal B}({\mathcal X}_1,{\mathcal X}_2)$, will be used for bounded operators between two Banach spaces ${\mathcal X}_1$ and ${\mathcal X}_2$. Moreover, ${\mathcal X}_1\hookrightarrow {\mathcal X}_2$ denotes the continuous embedding of ${\mathcal X}_1$ into ${\mathcal X}_2$. \section{General Boundary Value Problems and Boundary Data Maps} \label{s2} This section is devoted to boundary data maps and their basic properties. Taking $R>0$, and fixing $\theta_0, \theta_R\in S_{2 \pi}$, with $S_{2 \pi}$ the strip \begin{equation} S_{2 \pi}=\{z\in{\mathbb{C}}\,|\, 0\leq \Re(z) < 2 \pi\}, \end{equation} we introduce the linear map $\ensuremath{\ga_{\te_0,\te_R}}$, the trace map associated with the boundary $\{0,R\}$ of $(0,R)$ and the parameters $\theta_0, \theta_R$, by \begin{equation}\label{2.2} \ensuremath{\ga_{\te_0,\te_R}} \colon \begin{cases} C^1({[0,R]}) \rightarrow {\mathbb{C}}^2, \\ u \mapsto \begin{bmatrix} \cos(\theta_0)u(0) + \sin(\theta_0)u'(0)\\ \cos(\theta_R)u(R) - \sin(\theta_R)u'(R) \end{bmatrix}, \end{cases} \quad \theta_0, \theta_R\in S_{2 \pi}, \end{equation} where ``prime'' denotes $d/dx$. We note, in particular, that the Dirichlet trace $\gamma_D$, and the Neumnann trace $\gamma_N$ (in connection with the outward pointing unit normal vector at $\partial (0,R) = \{0, R\}$), are given by \begin{equation} \gamma_D = \gamma_{0,0} = - \gamma_{\pi,\pi}, \quad \gamma_N = \gamma_{3\pi/2,3\pi/2} = - \gamma_{\pi/2,\pi/2}. \label{2.2AA} \end{equation} Next, assuming \begin{equation} V\in L^1((0,R); dx), \label{2.2aa} \end{equation} we introduce the following family of densely defined closed linear operators $\ensuremath{H_{\te_0,\te_R}}$ in $L^2((0,R); dx)$, \begin{align} & \ensuremath{H_{\te_0,\te_R}} f= -f'' + Vf, \quad \theta_0, \theta_R\in S_{2 \pi}, \notag \\ & f\in\text{\rm{dom}}(\ensuremath{H_{\te_0,\te_R}})=\big\{ g \in L^2((0,R); dx)\,\big|\, g, g'\in AC({[0,R]}); \, \ensuremath{\ga_{\te_0,\te_R}} (g)=0; \label{2.2a} \\ & \hspace*{6.75cm} (-g''+Vg)\in L^2((0,R); dx)\big\}. \notag \end{align} Here $AC([0,R])$ denotes the set of absolutely continuous functions on $[0,R]$. We emphasize that $V$ is not assumed to be real-valued in the bulk of this paper. One notices that \begin{equation} \gamma_{(\theta_0 + \pi) \, \text{\rm mod} (2\pi), (\theta_R + \pi) \, \text{\rm mod} (2\pi)} = - \ensuremath{\ga_{\te_0,\te_R}}, \quad \theta_0, \theta_R\in S_{2 \pi}, \label{2.2A} \end{equation} and, on the other hand, \begin{equation} H_{(\theta_0 + \pi) \, \text{\rm mod} (2\pi), (\theta_R + \pi) \, \text{\rm mod} (2\pi)} = \ensuremath{H_{\te_0,\te_R}}, \quad \theta_0, \theta_R\in S_{2 \pi}, \label{2.2B} \end{equation} hence it suffices to consider $\theta_0, \theta_R\in S_{\pi}=\{z\in{\mathbb{C}}\,|\, 0\leq \Re(z) < \pi\}$ rather than $\theta_0, \theta_R\in S_{2 \pi}$ in connection with $\ensuremath{H_{\te_0,\te_R}}$, but for simplicity of notation we will keep using the strip $S_{2 \pi}$ throughout this manuscript. That $\ensuremath{H_{\te_0,\te_R}}$ is indeed a closed operator follows, for instance, from \cite[Sect.\ XII.4]{DS88}, especially, by combining Lemma 5\,(c) and the first part of the proof of Lemma 26 and noting that $g(0), g'(0)$ (resp., $g(R), g'(R)$) are a complete set of boundary values for the minimal operator $H_{\rm min}$ associated with the differential expression $-d^2/dx^2 + V(x)$ in $L^2((0,R); dx)$ at $x=0$ (resp., at $x=R$). Here \begin{align} & H_{\rm min} f= -f'' + Vf, \notag \\ & f\in\text{\rm{dom}}(H_{\rm min})=\big\{ g \in L^2((0,R); dx)\,\big|\, g, g'\in AC({[0,R]}); \\ & \hspace*{1cm} g(0)=g'(0)=g(R)=g'(R)=0; \, (-g''+Vg)\in L^2((0,R); dx)\big\}. \notag \end{align} Morever, the adjoint of $\ensuremath{H_{\te_0,\te_R}}$ is given by \begin{align} & (\ensuremath{H_{\te_0,\te_R}})^* f= -f'' + {\overline V} f, \quad \theta_0, \theta_R\in S_{2 \pi}, \notag \\ & f\in\text{\rm{dom}}\big((\ensuremath{H_{\te_0,\te_R}})^*\big)=\big\{ g \in L^2((0,R); dx)\,\big|\, g, g'\in AC({[0,R]}); \, \gamma_{\overline{\theta_0},\overline{\theta_R}} (g)=0; \notag \\ & \hspace*{6.2cm}(-g''+{\overline V}g)\in L^2((0,R); dx)\big\}. \end{align} The fact that the spectrum of $\ensuremath{H_{\te_0,\te_R}}$, $\sigma (\ensuremath{H_{\te_0,\te_R}})$, is discrete is well-known, but due to its importance in the context of this paper, we now briefly recall its proof following an argument in Marchenko \cite{Ma86}: \begin{lemma} [See, \cite{Ma86}, Sect.\ 1.3] \label{l2.1} Suppose $V\in L^1((0,R); dx)$, assume $\theta_0, \theta_R\in S_{2 \pi}$, and let $\ensuremath{H_{\te_0,\te_R}}$ be defined as in \eqref{2.2a}. Then $\sigma (\ensuremath{H_{\te_0,\te_R}})$ is an infinite discrete subset of ${\mathbb{C}}$ $($i.e., a set without any finite limit point in ${\mathbb{C}}$, but with a limit point at infinity$)$. \end{lemma} \begin{proof} Fix $z\in{\mathbb{C}}$. Let $\theta(z,\cdot), \theta^{\prime}(z,\cdot), \phi(z,\cdot), \phi^{\prime}(z,\cdot)\in AC([0,R])$, and such that $\theta (z,\cdot)$ and $\phi (z,\cdot)$ are solutions of $-\psi'' +V \psi=z \psi$ uniquely determined by the initial values at $x=0$, \begin{equation} \theta (z,0)=\phi^{\prime}(z,0)=1, \quad \theta^{\prime}(z,0)=\phi (z,0)=0. \label{2.2b} \end{equation} Consequently, $\theta (z,\cdot)$ and $\phi (z,\cdot)$ are entire with respect to $z$. Introducing \begin{equation} \psi(z,\cdot) = A \theta (z,\cdot) + B \phi (z,\cdot), \quad A, B \in {\mathbb{C}}, \end{equation} it follows that \begin{equation} \ensuremath{\ga_{\te_0,\te_R}}(\psi) = \begin{bmatrix} \cos(\theta_0) \psi(z,0) + \sin(\theta_0) \psi'(z,0) \\[1mm] \cos(\theta_R) \psi(z,R) - \sin(\theta_R) \psi'(z,R) \end{bmatrix} = 0 \in {\mathbb{C}}^2. \label{2.2c} \end{equation} Employing the initial conditions \eqref{2.2b}, one concludes that equation \eqref{2.2c} is equivalent to \begin{align} 0 &= \begin{bmatrix} \cos(\theta_0) & \sin(\theta_0) \\[1mm] \cos(\theta_R) \theta (z,R) - \sin(\theta_R) \theta' (z,R) & \cos(\theta_R) \phi(z,R) - \sin(\theta_R) \phi'(z,R) \end{bmatrix} \begin{bmatrix} A \\[1mm] B \end{bmatrix} \notag \\ & = {\mathcal U}(z, R, \theta_0, \theta_R) \begin{bmatrix} A \\ B \end{bmatrix}. \label{2.2ca} \end{align} Consequently, $z_0$ is an eigenvalue of $\ensuremath{H_{\te_0,\te_R}}$ if and only if $z_0$ is a zero of the determinant $\Delta$ defined as \begin{equation} \Delta(z,R,\theta_0,\theta_R) =\det\big({\mathcal U}(z, R, \theta_0, \theta_R)\big). \label{2.2det} \end{equation} Thus, $\Delta$ is an entire function with respect to $z$, and an explicit computation reveals that \begin{align} \begin{split} \Delta(z,R,\theta_0,\theta_R)&=\cos(\theta_0)\cos(\theta_R)\phi (z,R) - \cos(\theta_0)\sin(\theta_R) \phi^{\prime} (z,R) \\ & \quad - \sin(\theta_0)\cos(\theta_R) \theta (z,R) + \sin(\theta_0)\sin(\theta_R) \theta^{\prime} (z,R). \label{2.2d} \end{split} \end{align} The standard Volterra integral equations \begin{align} \theta (z,x) &= \cos(z^{1/2}x) + \int_0^x dx' \, \frac{\sin(z^{1/2}(x-x'))}{z^{1/2}} V(x') \theta (z,x'), \\ \phi (z,x) &= \frac{\sin(z^{1/2}x)}{z^{1/2}} + \int_0^x dx' \, \frac{\sin(z^{1/2}(x-x'))}{z^{1/2}} V(x') \phi (z,x'), \\ & \hspace*{3.5cm} z\in{\mathbb{C}}, \; \Im(z^{1/2}) \geq 0, \; x\in [0,R], \notag \end{align} then imply that \begin{align} \begin{split} \theta (z,x) & \underset{|z|\to\infty}{=} \cos(z^{1/2}x) + O\Big(|z|^{-1/2}e^{\Im(z^{1/2})x}\Big), \\ \theta^{\prime}(z,x) & \underset{|z|\to\infty}{=} - z^{1/2} \sin(z^{1/2}x) + O\Big(e^{\Im(z^{1/2})x}\Big), \\ \phi (z,x) & \underset{|z|\to\infty}{=} \frac{\sin(z^{1/2}x)}{z^{1/2}} + O\Big(|z|^{-1}e^{\Im(z^{1/2})x}\Big), \label{2.2e} \\ \phi^{\prime}(z,x) & \underset{|z|\to\infty}{=} \cos(z^{1/2}x) + O\Big(|z|^{-1/2}e^{\Im(z^{1/2})x}\Big). \end{split} \end{align} A comparison of \eqref{2.2d} as $|z|\to\infty$ and \eqref{2.2e} demonstrates that $\Delta$ does not vanish identically. Thus the set of zeros of $\Delta$, and hence the set of eigenvalues of $\ensuremath{H_{\te_0,\te_R}}$, constitutes a discrete set. Again, the asymptotic behavior of $\Delta$ near infinity implies that $\Delta$ is an entire function of order $1/2$ and hence possesses infinitely many zeros (cf., e.g., \cite[p.\ 252]{Ti85}). \end{proof} In addition (cf.\ also \eqref{3.32}, \eqref{3.33}), the resolvent of $\ensuremath{H_{\te_0,\te_R}}$ is clearly a Hilbert--Schmidt operator in $L^2((0,R); dx)$. In fact, it is even a trace class operator since the eigenvalues $E_{\theta_0,\theta_R,n}$ of $\ensuremath{H_{\te_0,\te_R}}$ in the case of the separated boundary conditions at hand are of the form $E_{\theta_0,\theta_R,n} = [(n\pi/R) + (a_n/n)]^2$ with $a_n\in \ell^\infty({\mathbb{N}})$ as $n\to \infty$, as shown in \cite[Lemma 1.3.3]{Ma86}. Having described the operator $\ensuremath{H_{\te_0,\te_R}}$ is some detail, still assuming \eqref{2.2aa}, we now briefly recall the corresponding closed, sectorial, and densely defined sequilinear form, denoted by $Q_{\ensuremath{H_{\te_0,\te_R}}}$, associated with $\ensuremath{H_{\te_0,\te_R}}$ (cf.\ \cite[p.\ 312, 321, 327--328]{Ka80}): \begin{align} & Q_{\ensuremath{H_{\te_0,\te_R}}}(f,g) = \int_0^R dx \big[\overline{f'(x)} g'(x) + V(x) \overline{f(x)} g(x)\big] \notag \\ & \hspace*{2.35cm} - \cot(\theta_0) \overline{f(0)} g(0) - \cot(\theta_R) \overline{f(R)} g(R), \label{2.2f} \\ & f, g \in \text{\rm{dom}}(Q_{\ensuremath{H_{\te_0,\te_R}}}) = \text{\rm{dom}}\big(|\ensuremath{H_{\te_0,\te_R}}|^{1/2}\big) = H^1((0,R)) \notag \\ & \quad = \big\{h\in L^2((0,R); dx) \,|\, h \in AC ([0,R]); \, h' \in L^2((0,R); dx)\big\}, \notag \\ & \hspace*{6.5cm} \theta_0, \theta_R \in S_{2 \pi}\backslash\{0,\pi\}, \notag \\ & Q_{H_{0,\theta_R}}(f,g) = \int_0^R dx \big[\overline{f'(x)} g'(x) + V(x) \overline{f(x)} g(x)\big] - \cot(\theta_R) \overline{f(R)} g(R), \label{2.2g} \\ & f, g \in \text{\rm{dom}}(Q_{H_{0,\theta_R}}) = \text{\rm{dom}}\big(|H_{0,\theta_R}|^{1/2}\big) \notag \\ & \quad = \big\{h\in L^2((0,R); dx) \,|\, h \in AC ([0,R]); \, h(0) =0; \, h' \in L^2((0,R); dx)\big\}, \notag \\ & \hspace*{8.5cm} \theta_R \in S_{2 \pi}\backslash\{0,\pi\}, \notag \\ & Q_{H_{\theta_0,0}}(f,g) = \int_0^R dx \big[\overline{f'(x)} g'(x) + V(x) \overline{f(x)} g(x)\big] - \cot(\theta_0) \overline{f(0)} g(0), \label{2.2h} \\ & f, g \in \text{\rm{dom}}(Q_{H_{\theta_0,0}}) = \text{\rm{dom}}\big(|H_{\theta_0,0}|^{1/2}\big) \notag \\ & \quad = \big\{h\in L^2((0,R); dx) \,|\, h \in AC ([0,R]); \, h(R) =0; \, h' \in L^2((0,R); dx)\big\}, \notag \\ & \hspace*{8.67cm} \theta_0 \in S_{2 \pi}\backslash\{0,\pi\}, \notag \\ & Q_{H_{0,0}}(f,g) = \int_0^R dx \big[\overline{f'(x)} g'(x) + V(x) \overline{f(x)} g(x)\big], \label{2.2i} \\ & f, g \in \text{\rm{dom}}(Q_{H_{0,0}}) = \text{\rm{dom}}\big(|H_{0,0}|^{1/2}\big) = H^1_0((0,R)) \notag \\ & \quad = \big\{h\in L^2((0,R); dx) \,|\, h \in AC([0,R]); \, h(0)=0, \, h(R) =0; \notag \\ & \hspace*{6.65cm} h' \in L^2((0,R); dx)\big\}. \notag \end{align} Equations \eqref{2.2f}--\eqref{2.2i} follow from the fact that for any $\varepsilon>0$, there exists $\eta(\varepsilon)>0$ such that for all $h \in H^1((0,R))$, \begin{align} & |h(x_0)| \leq \varepsilon \|h' \|_{L^2((0,R); dx)} + \eta(\varepsilon) \|h \|_{L^2((0,R); dx)}, \quad x_0 \in [0,R], \\ & \||V|^{1/2} h\|_{L^2((0,R); dx)} \leq \varepsilon \|h' \|_{L^2((0,R); dx)} + \eta(\varepsilon) \|h \|_{L^2((0,R); dx)} \end{align} (cf.\ \cite[p.\ 193, 345--346]{Ka80}). Next, we recall the following elementary, yet fundamental, fact: \begin{lemma} \label{l2.2} Suppose that $V\in L^1((0,R); dx)$, fix $\theta_0, \theta_R\in S_{2 \pi}$, and assume that $z\in{\bbC\backslash\sigma(\ensuremath{H_{\te_0,\te_R}})}$. Then the boundary value problem given by \begin{align -&u'' + Vu=zu,\quad u, u'\in AC([0,R]), \label{2.3}\\ &\ensuremath{\ga_{\te_0,\te_R}}(u)=\begin{bmatrix}c_0\\ c_R \end{bmatrix}\in{\mathbb{C}}^2, \label{2.4} \end{align} has a unique solution $u(z,\cdot)=u(z,\cdot\,;(\theta_0,c_0),(\theta_R,c_R))$ for each $c_0, c_R\in{\mathbb{C}}$. \end{lemma} \begin{proof} This is well-known, but for the sake of completeness, we briefly recall the argument: Let $\psi_j(z,\cdot)$, $j=1,2$, be a basis for the solutions of \eqref{2.3} and let $\psi(z,\cdot)=A\psi_1(z,\cdot)+B\psi_2(z,\cdot)$, $A,B\in{\mathbb{C}}$, be the general solution of \eqref{2.3}. Then \begin{align} \begin{split} \ensuremath{\ga_{\te_0,\te_R}}(\psi(z,\cdot)) &= \begin{bmatrix}\ensuremath{\ga_{\te_0,\te_R}}(\psi_1(z,\cdot)) & \ensuremath{\ga_{\te_0,\te_R}}(\psi_2(z,\cdot)) \end{bmatrix} \begin{bmatrix} A \\ B \end{bmatrix} \\ & = \begin{bmatrix} M_{1,1}(z) & M_{1,2}(z) \\ M_{2,1}(z) & M_{2,2}(z) \end{bmatrix} \begin{bmatrix} A\\ B \end{bmatrix}, \label{2.4a} \end{split} \end{align} where \begin{align} \begin{split} M_{1,1}(z) &= \cos(\theta_0)\psi_1(z,0) + \sin(\theta_0)\psi_1'(z,0), \\ M_{1,2}(z) &= \cos(\theta_0)\psi_2(z,0) + \sin(\theta_0)\psi_2'(z,0), \\ M_{2,1}(z) &= \cos(\theta_R)\psi_1(z,R) - \sin(\theta_R)\psi_1'(z,R), \\ M_{2,2}(z) &= \cos(\theta_R)\psi_2(z,R) - \sin(\theta_R)\psi_2'(z,R). \end{split} \end{align} Thus, prescribing $c_0, c_R \in{\mathbb{C}}$, the equation \begin{equation} \ensuremath{\ga_{\te_0,\te_R}}(\psi(z,\cdot))=\begin{bmatrix}c_0\\ c_R \end{bmatrix} \end{equation} is uniquely solvable in terms of some $A,B\in{\mathbb{C}}$ if and only if \begin{align} \det\big(\begin{bmatrix}\ensuremath{\ga_{\te_0,\te_R}}(\psi_1(z,\cdot)) & \ensuremath{\ga_{\te_0,\te_R}}(\psi_2(z,\cdot))\end{bmatrix}\big) =\det\left(\begin{bmatrix} M_{1,1}(z) & M_{1,2}(z) \\ M_{2,1}(z) & M_{2,2}(z) \end{bmatrix} \right) \ne 0. \label{2.10} \end{align} On the other hand, this determinant equals zero for some $z_0\in{\mathbb{C}}$, if and only if there is a nonzero vector $\begin{bmatrix} A_0 & B_0 \end{bmatrix}^\top \in{\mathbb{C}}^2$ such that \begin{equation} \begin{bmatrix}\ensuremath{\ga_{\te_0,\te_R}}(\psi_1(z_0,\cdot)) & \ensuremath{\ga_{\te_0,\te_R}}(\psi_2(z_0,\cdot))\end{bmatrix} \begin{bmatrix} A_0 \\ B_0 \end{bmatrix} =0 \end{equation} which is equivalent to the existence of a nonzero solution $\psi_0(z_0,\cdot)= A_0 \psi_1(z_0,\cdot) + B_0 \psi_2(z_0,\cdot)$ of the corresponding boundary value problem given by \eqref{2.3} and \eqref{2.4} with $z=z_0$ and homogeneous boundary conditions (i.e., with $c_0=c_R=0$). Equivalently, $\psi_0(z_0,\cdot)$ satisfies \begin{equation} \ensuremath{H_{\te_0,\te_R}} \psi_0(z_0,\cdot) = z_0 \psi_0(z_0,\cdot), \quad \psi_0(z_0,\cdot) \in \text{\rm{dom}}(\ensuremath{H_{\te_0,\te_R}}), \end{equation} which in turn is equivalent to $z_0\in\sigma(\ensuremath{H_{\te_0,\te_R}})$. \end{proof} Assuming $z\in{\mathbb{C}}\backslash\sigma(\ensuremath{H_{\te_0,\te_R}})$, a basis for the solutions of \eqref{2.3} is given by \begin{equation}\label{2.5} \begin{split} u_{-,\theta_0}(z,\cdot)&=u(z,\cdot \, ;(\theta_0,0),(0,1)), \\ u_{+,\theta_R}(z,\cdot)&=u(z,\cdot \, ;(0,1),(\theta_R,0)). \end{split} \end{equation} Explicitly, one then has \begin{align} u_{-,\theta_0}(z,R)&= 1, \quad \cos(\theta_0)u_{-,\theta_0}(z,0) + \sin(\theta_0)u_{-,\theta_0}'(z,0) =0, \label{2.5a}\\ u_{+,\theta_R}(z,0)&= 1, \quad \cos(\theta_R)u_{+,\theta_R}(z,R) - \sin(\theta_R)u_{+,\theta_R}'(z,R) =0. \label{2.5b} \end{align} Recalling the Wronskian of two functions $f$ and $g$, \begin{equation} W(f,g)(x) = f(x)g'(x) - f'(x)g(x), \quad f, g \in C^1([0,R]), \end{equation} one then computes \begin{align} & W(u_{+,\theta_R}(z,\cdot), u_{-,\theta_0}(z,\cdot)) \notag \\ & \quad = u_{+,\theta_R}(z,x) u'_{-,\theta_0}(z,x)- u'_{+,\theta_R}(z,x) u_{-,\theta_0}(z,x) \neq 0, \quad x \in [0,R], \notag \\ & \quad = u'_{-,\theta_0}(z,0)- u'_{+,\theta_R}(z,0), u_{-,\theta_0}(z,0) \label{2.5c} \\ & \quad = u_{+,\theta_R}(z,R) u'_{-,\theta_0}(z,R)- u'_{+,\theta_R}(z,R). \label{2.5d} \end{align} To each boundary value problem \eqref{2.3}, \eqref{2.4}, we now associate a family of \emph{general boundary data maps}, $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z) : {\mathbb{C}}^2 \rightarrow {\mathbb{C}}^2$, for $\theta_0, \theta_R, \theta_0^{\prime},\theta_R^{\prime}\in S_{2 \pi}$, where \begin{align}\label{2.6} \begin{split} \ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z) \begin{bmatrix}c_0\\ c_R \end{bmatrix} &= \ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z) \big(\ensuremath{\ga_{\te_0,\te_R}}(u(z,\cdot\ ;(\theta_0,c_0),(\theta_R,c_R)))\big) \\ &= \ensuremath{\ga_{\te_0^{\prime},\te_R^{\prime}}}(u(z,\cdot\ ;(\theta_0,c_0),(\theta_R,c_R))). \end{split} \end{align} With $u(z,\cdot)=u(z,\cdot\ ;(\theta_0,c_0),(\theta_R,c_R))$, then $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z) $ can be represented as a $2\times 2$ complex matrix, where \begin{align}\label{2.7} \ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z) \begin{bmatrix}c_0\\ c_R \end{bmatrix} &= \ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z) \begin{bmatrix} \cos(\theta_0)u(z,0) + \sin(\theta_0)u'(z,0)\\[1mm] \cos(\theta_R)u(z,R) - \sin(\theta_R)u'(z,R) \end{bmatrix} \notag \\ &= \begin{bmatrix} \cos(\theta_0^{\prime})u(z,0) + \sin(\theta_0^{\prime})u'(z,0)\\[1mm] \cos(\theta_R^{\prime})u(z,R) - \sin(\theta_R^{\prime})u'(z,R) \end{bmatrix}. \end{align} The following result shows that $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}}$ is well-defined for $z\in{\mathbb{C}}\backslash\sigma(\ensuremath{H_{\te_0,\te_R}})$, that is, it is invariant with respect to a change of basis of solutions of \eqref{2.3}. \begin{theorem}\label{t2.3} Let $\theta_0, \theta_R, \theta_0^{\prime},\theta_R^{\prime}\in S_{2 \pi}$ and $z\in{\mathbb{C}}\backslash\sigma(\ensuremath{H_{\te_0,\te_R}})$. In addition, denote by $\psi_j(z,\cdot)$, $j=1,2$, a basis for the solutions of \eqref{2.3}. Then, \begin{align}\label{2.9} & \ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z) \\ & \;\; =\begin{bmatrix}\cos(\theta_0^{\prime})\psi_1(z,0) + \sin(\theta_0^{\prime})\psi_1'(z,0)&\cos(\theta_0^{\prime})\psi_2(z,0) + \sin(\theta_0^{\prime})\psi_2'(z,0)\\[1mm] \cos(\theta_R^{\prime})\psi_1(z,R) - \sin(\theta_R^{\prime})\psi_1'(z,R)&\cos(\theta_R^{\prime})\psi_2(z,R) - \sin(\theta_R^{\prime})\psi_2'(z,R)\end{bmatrix} \notag \\ & \;\;\;\; \times \begin{bmatrix}\cos(\theta_0)\psi_1(z,0) + \sin(\theta_0)\psi_1'(z,0)&\cos(\theta_0)\psi_2(z,0) + \sin(\theta_0)\psi_2'(z,0)\\[1mm] \cos(\theta_R)\psi_1(z,R) - \sin(\theta_R)\psi_1'(z,R)&\cos(\theta_R)\psi_2(z,R) - \sin(\theta_R)\psi_2'(z,R)\end{bmatrix}^{-1}. \notag \end{align} Moreover, $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z) $ is invariant with respect to a change of basis for the solutions of \eqref{2.3}. \end{theorem} \begin{proof} Letting $\psi(z,\cdot)=A\psi_1(z,\cdot)+B\psi_2(z,\cdot)$, $A,B\in{\mathbb{C}}$, be an arbitrary solution of \eqref{2.3}, one observes, by \eqref{2.4a} and \eqref{2.7}, that the equation $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}}(\ensuremath{\ga_{\te_0,\te_R}}(\psi))=\ensuremath{\ga_{\te_0^{\prime},\te_R^{\prime}}}(\psi)$ becomes \begin{align}\label{2.13} & \ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}}(z) \\ & \; \times \begin{bmatrix}\cos(\theta_0)\psi_1(z,0) + \sin(\theta_0)\psi_1'(z,0)&\cos(\theta_0)\psi_2(z,0) + \sin(\theta_0)\psi_2'(z,0)\\[1mm] \cos(\theta_R)\psi_1(z,R) - \sin(\theta_R)\psi_1'(z,R)&\cos(\theta_R)\psi_2(z,R) - \sin(\theta_R)\psi_2'(z,R)\end{bmatrix}\begin{bmatrix}A\\B\end{bmatrix} \notag \\ & \;\; =\begin{bmatrix}\cos(\theta_0^{\prime})\psi_1(z,0) + \sin(\theta_0^{\prime})\psi_1'(z,0)&\cos(\theta_0^{\prime})\psi_2(z,0) + \sin(\theta_0^{\prime})\psi_2'(z,0)\\[1mm] \cos(\theta_R^{\prime})\psi_1(z,R) - \sin(\theta_R^{\prime})\psi_1'(z,R)&\cos(\theta_R^{\prime})\psi_2(z,R) - \sin(\theta_R^{\prime})\psi_2'(z,R)\end{bmatrix}\begin{bmatrix}A\\B\end{bmatrix} \notag \end{align} for every $\begin{bmatrix}A & B\end{bmatrix}^\top \in{\mathbb{C}}^2$. Equation \eqref{2.9} then follows by the invertibility of $\begin{bmatrix}\ensuremath{\ga_{\te_0,\te_R}}(\psi_1) & \ensuremath{\ga_{\te_0,\te_R}}(\psi_2)\end{bmatrix}$ noted in \eqref{2.10}. Let $\phi_j(z,\cdot)$, $j=1,2$, denote a second basis for the solutions of \eqref{2.3}. Then, there is a nonsingular matrix $K\in{\mathbb{C}}^{2\times 2}$ such that $\begin{bmatrix}\psi_1 & \psi_2\end{bmatrix}=\begin{bmatrix}\phi_1 & \phi_2 \end{bmatrix}K$. Next, we introduce for each pair, $\theta_0, \theta_R\in S_{2 \pi}$, the following matrices \begin{equation}\label{2.14} \ensuremath{C_{\te_0,\te_R}}=\begin{bmatrix}\cos(\theta_0)&0\\0&\cos(\theta_R) \end{bmatrix},\quad \ensuremath{S_{\te_0,\te_R}}=\begin{bmatrix} \sin(\theta_0)&0\\0&\sin(\theta_R)\end{bmatrix}. \end{equation} Introducing $\theta_j$ to denote $\theta_0$, and $\theta_k$ to denote $\theta_R$, respectively; or, using $\theta_j$ to denote $\theta_0^{\prime}$ and $\theta_k$ to denote $\theta_R^{\prime}$, one computes, \begin{align} & \begin{bmatrix}\gamma_{\theta_j,\theta_k}(\psi_1(z,\cdot)) & \gamma_{\theta_j,\theta_k}(\psi_2(z,\cdot))\end{bmatrix} \notag \\ & \quad = C_{\theta_j,\theta_k}\begin{bmatrix} \psi_1(z,0)&\psi_2(z,0)\\ \psi_1(z,R)&\psi_2(z,R)\end{bmatrix} + S_{\theta_j,\theta_k}\begin{bmatrix} \psi'_1(z,0)&\psi'_2(z,0)\\ -\psi'_1(z,R)&-\psi'_2(z,R)\end{bmatrix} \notag \\ & \quad = \bigg(C_{\theta_j,\theta_k}\begin{bmatrix} \phi_1(z,0)&\phi_2(z,0)\\ \phi_1(z,R)&\phi_2(z,R)\end{bmatrix} + S_{\theta_j,\theta_k}\begin{bmatrix} \phi'_1(z,0)&\phi'_2(z,0)\\ -\phi'_1(z,R)&-\phi'_2(z,R)\end{bmatrix}\bigg)K \notag \\ & \quad = \begin{bmatrix}\gamma_{\theta_j,\theta_k}(\phi_1(z,\cdot)) & \gamma_{\theta_j,\theta_k}(\phi_2(z,\cdot)) \end{bmatrix} K. \label{2.16} \end{align} As defined in \eqref{2.9}, $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}}=\begin{bmatrix}\gamma_{\theta_0^{\prime},\theta_R^{\prime}}(\psi_1) & \gamma_{\theta_0^{\prime},\theta_R^{\prime}}(\psi_2)\end{bmatrix}\begin{bmatrix} \gamma_{\theta_0,\theta_R}(\psi_1) & \gamma_{\theta_0,\theta_R}(\psi_2)\end{bmatrix}^{-1}$, and by \eqref{2.16}, \begin{align}\label{2.32} \begin{split} \ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} & = \begin{bmatrix}\gamma_{\theta_0^{\prime},\theta_R^{\prime}}(\psi_1) & \gamma_{\theta_0^{\prime},\theta_R^{\prime}}(\psi_2)\end{bmatrix}\begin{bmatrix}\gamma_{\theta_0,\theta_R}(\psi_1) &\gamma_{\theta_0,\theta_R}(\psi_2)\end{bmatrix}^{-1} \\ & =\begin{bmatrix}\gamma_{\theta_0^{\prime},\theta_R^{\prime}}(\phi_1) & \gamma_{\theta_0^{\prime},\theta_R^{\prime}}(\phi_2)\end{bmatrix}\begin{bmatrix}[\gamma_{\theta_0,\theta_R}(\phi_1) & \gamma_{\theta_0,\theta_R}(\phi_2)\end{bmatrix}^{-1}, \end{split} \end{align} completing the proof. \end{proof} Theorem \ref{t2.3} then readily implies the following result: \begin{corollary} \label{c2.4} Let $\theta_0,\theta_R, \theta_0^{\prime},\theta_R^{\prime}, \theta_0^{\prime\prime}, \theta_R^{\prime\prime}\in S_{2 \pi}$. Then, with $I_2$ denoting the identity matrix in ${\mathbb{C}}^2$, \begin{align & \Lambda_{\theta_0,\theta_R}^{\theta_0,\theta_R} (z) = I_2, \quad z\in{\mathbb{C}} \backslash \sigma(\ensuremath{H_{\te_0,\te_R}}), \\ & \Lambda_{\theta_0^{\prime},\theta_R^{\prime}}^{\theta_0^{\prime\prime},\theta_R^{\prime\prime}} (z) \ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z) =\Lambda_{\theta_0,\theta_R}^{\theta_0^{\prime\prime},\theta_R^{\prime\prime}} (z) , \quad z\in{\mathbb{C}}\backslash\big(\sigma(\ensuremath{H_{\te_0,\te_R}})\cup\sigma(H_{\theta'_0,\theta'_R})\big), \\ & \Lambda_{\theta_0^{\prime},\theta_R^{\prime}}^{\theta_0,\theta_R} (z) = \Big[\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z)\Big]^{-1}, \quad z\in{\mathbb{C}}\backslash\big(\sigma(\ensuremath{H_{\te_0,\te_R}})\cup\sigma(H_{\theta'_0,\theta'_R})\big). \label{2.48} \end{align} \end{corollary} \begin{remark} \label{r2.5} By Theorem \ref{t2.3}, $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}}$ is invariant with respect to a change of basis for the solutions of \eqref{2.3}. However, the representation of $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}}$ with respect to a specific basis can be simplified considerably with an appropriate choice of basis. For example, by choosing the basis given in \eqref{2.5}, and by letting $\psi_1(z,\cdot)=u_{+,\theta_R}(z,\cdot)=u(z,\cdot \, ;(0,1),(\theta_R,0))$, and $\psi_2(z,\cdot)=u_{-,\theta_0}(z,\cdot)=u(z,\cdot \, ;(\theta_0,0),(0,1))$, \eqref{2.32} implies that, entrywise, \begin{align \begin{split} & \begin{bmatrix}\ensuremath{\ga_{\te_0,\te_R}}(u_{+,\theta_R}(z,\cdot)) & \ensuremath{\ga_{\te_0,\te_R}}(u_{-,\theta_0}(z,\cdot))\end{bmatrix}_{1,1} \\ & \quad = \cos(\theta_0) + \sin(\theta_0)u_{+,\theta_R}'(z,0), \\ & \begin{bmatrix}\ensuremath{\ga_{\te_0,\te_R}}(u_{+,\theta_R}(z,\cdot)) & \ensuremath{\ga_{\te_0,\te_R}}(u_{-,\theta_0}(z,\cdot))\end{bmatrix}_{1,2} \\ & \quad = \cos(\theta_0)u_{-,\theta_0}(z,0) + \sin(\theta_0)u_{-,\theta_0}'(z,0), \\ & \begin{bmatrix}\ensuremath{\ga_{\te_0,\te_R}}(u_{+,\theta_R}(z,\cdot)) & \ensuremath{\ga_{\te_0,\te_R}}(u_{-,\theta_0}(z,\cdot))\end{bmatrix}_{2,1} \\ & \quad = \cos(\theta_R)u_{+,\theta_R}(z,R) - \sin(\theta_R)u_{+,\theta_R}'(z,R), \\ & \begin{bmatrix}\ensuremath{\ga_{\te_0,\te_R}}(u_{+,\theta_R}(z,\cdot)) & \ensuremath{\ga_{\te_0,\te_R}}(u_{-,\theta_0}(z,\cdot))\end{bmatrix}_{2,1} \\ & \quad = \cos(\theta_R) - \sin(\theta_R)u_{-,\theta_0}'(z,R). \end{split} \end{align} Hence, for this basis, \begin{align} & \ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z) = \Big[\Big(\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z)\Big)_{j,k}\Big]_{1 \leq j,k \leq 2}, \quad z\in{\mathbb{C}}\backslash\sigma(\ensuremath{H_{\te_0,\te_R}}), \label{2.20} \\[1mm] \begin{split} & \Big(\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z)\Big)_{1,1} = \frac{\cos(\theta_0^{\prime}) + \sin(\theta_0^{\prime})u_{+,\theta_R}'(z,0)}{\cos(\theta_0) + \sin(\theta_0)u_{+,\theta_R}'(z,0)}, \\ & \Big(\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z)\Big)_{1,2} = \frac{\cos(\theta_0^{\prime})u_{-,\theta_0}(z,0) + \sin(\theta_0^{\prime})u_{-,\theta_0}'(z,0)}{\cos(\theta_R) - \sin(\theta_R)u_{-,\theta_0}'(z,R)}, \\ & \Big(\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z)\Big)_{2,1} = \frac{\cos(\theta_R^{\prime})u_{+,\theta_R}(z,R) - \sin(\theta_R^{\prime})u_{+,\theta_R}'(z,R)}{\cos(\theta_0) + \sin(\theta_0)u_{+,\theta_R}'(z,0)}, \\ & \Big(\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z)\Big)_{2,2} = \frac{\cos(\theta_R^{\prime}) - \sin(\theta_R^{\prime})u_{-,\theta_0}'(z,R)}{\cos(\theta_R) - \sin(\theta_R)u_{-,\theta_0}'(z,R)}. \label{2.20a} \end{split} \end{align} In particular, by \eqref{2.5a} and \eqref{2.5b}, \begin{equation} \Big(\Lambda_{\theta_0,\theta_R}^{\theta_0,\theta_R'} (z)\Big)_{1,2} = 0, \quad \Big(\Lambda_{\theta_0,\theta_R}^{\theta_0',\theta_R} (z)\Big)_{2,1} = 0. \end{equation} \end{remark} \begin{remark} \label{r2.6} We note that $\ensuremath{\La_{0,0}^{\frac{\pi}{2},\frac{\pi}{2}}}(z)$ represents the \emph{Dirichlet-to-Neumann map}, $\Lambda_{D,N}(z)$, for the boundary value problem \eqref{2.3}, \eqref{2.4}; that is, when $\theta_0=\theta_R=0$, $\theta_0^{\prime}=\theta_R^{\prime}=\pi/2$, then \eqref{2.7} becomes \begin{equation \Lambda_{D,N}(z) \begin{bmatrix}u(z,0)\\u(z,R)\end{bmatrix}= \ensuremath{\La_{0,0}^{\frac{\pi}{2},\frac{\pi}{2}}}( z) \begin{bmatrix}u(z,0)\\u(z,R)\end{bmatrix}= \left[\negthickspace\begin{array}{r}u'(z,0)\\-u'(z,R)\end{array}\negthickspace\right], \quad z\in{\mathbb{C}}\backslash\sigma(\ensuremath{H_{\te_0,\te_R}}), \end{equation} with $u(z,\cdot)=u(z,\cdot \, ;(0,c_0),(0,c_R))$, $u(z,0)=c_0$, $u(z,R)=c_R$. The Dirichlet-to-Neumann map in the case $V=0$ has recently been considered in \cite[Example 5.1]{Po08}. The Neumann-to-Dirichlet map $\Lambda_{N,D}(z) = \Lambda_{\pi/2,\pi/2}^{\pi,\pi} (z) = - [\Lambda_{D,N}(z)]^{-1}$ (cf.\ \eqref{4.60}) in the case $V=0$ has earlier been computed in \cite[Example\ 4.1]{DM92}. We also refer to \cite{BMN08}, \cite{BGP08}, \cite{DM91}, \cite{FOP06} in the intimately related context of $Q$ and $M$-functions. \end{remark} It would be interesting to establish precise connections between $\Lambda_{\theta_0,\theta_R}^{\theta_0,\theta_R} (z)$ and the dynamical response operator discussed, for instance, in \cite{ABI92}, \cite{AK08}, \cite{ALP02}, \cite{ALP05} in connection with the problem of regularity and controllability of the wave equation on a compact interval. We conclude this section with an elementary result needed in the proof of Lemma \ref{l3.4} and Theorem \ref{t4.6}: \begin{lemma} \label{l2.7} Let $V\in L^1((0,R); dx)$, fix $\theta_0, \theta_R\in S_{2 \pi}$, $c_0, c_R\in{\mathbb{C}}$, and assume that $z\in{\bbC\backslash\sigma(\ensuremath{H_{\te_0,\te_R}})}$. Then the unique solution $u(z,\cdot)=u(z,\cdot\,;(\theta_0,c_0),(\theta_R,c_R))$ of \eqref{2.3}, \eqref{2.4} can be expressed in terms of the fundamental system $\theta (z,\cdot), \phi (z,\cdot)$, as introduced in the proof of Lemma \ref{l2.1}, and satisfying the initial conditions \eqref{2.2b}, as follows: \begin{align} & \, u(z,\cdot)=u(z,\cdot\,;(\theta_0,c_0),(\theta_R,c_R)) = \alpha (z) \theta (z,\cdot) + \beta (z) \phi (z,\cdot), \label{2.37} \\ & \begin{bmatrix} \alpha (z) \\[1mm] \beta (z) \end{bmatrix} = \frac{1}{\Delta (z,R,\theta_0,\theta_R)} \begin{bmatrix} [\cos(\theta_R) \phi (z,R) - \sin(\theta_R) \phi' (z,R)] c_0 - \sin(\theta_0) c_R \\[1mm] - [\cos(\theta_R) \theta (z,R) - \sin(\theta_R) \theta' (z,R)] c_0 + \cos(\theta_0) c_R \end{bmatrix}, \notag \end{align} with $\Delta$ given by \eqref{2.2d}. In particular, one has \begin{align} \begin{split} & \, u_{+,\theta_R}(z,\cdot)=u(z,\cdot\,;((0,1),(\theta_R,0)) = \alpha (z) \theta (z,\cdot) + \beta (z) \phi (z,\cdot), \label{2.38} \\ & \begin{bmatrix} \alpha (z) \\[1mm] \beta (z) \end{bmatrix} = \frac{1}{\Delta (z,R,0,\theta_R)} \begin{bmatrix} \cos(\theta_R) \phi (z,R) - \sin(\theta_R) \phi' (z,R) \\[1mm] - \cos(\theta_R) \theta (z,R) + \sin(\theta_R) \theta' (z,R) \end{bmatrix}, \end{split} \intertext{and} & \, u_{-,\theta_0}(z,\cdot)=u(z,\cdot\,;(\theta_0,0),(0,1)) \label{2.39} \\ & \, \qquad \qquad \, = \frac{1}{\Delta (z,R,\theta_0,0)} \big(-\sin(\theta_0) \theta (z,\cdot) + \cos(\theta_0) \phi (z,\cdot)\big). \notag \end{align} \end{lemma} \begin{proof} With $u(z,\cdot)=u(z,\cdot\,;(\theta_0,c_0),(\theta_R,c_R)) = \alpha \theta (z,\cdot) + \beta \phi (z,\cdot)$ one observes from \eqref{2.4} that \begin{align} & \begin{bmatrix} c_0 \\ c_R \end{bmatrix} = \gamma_{\theta_0,\theta_R} (u(z,\cdot)) = \gamma_{\theta_0,\theta_R} (\alpha \theta (z,\cdot) + \beta \phi (z,\cdot)) \notag \\ & \quad = \begin{bmatrix} \gamma_{\theta_0,\theta_R} (\theta (z,\cdot)) \quad \gamma_{\theta_0,\theta_R} (\phi (z,\cdot)) \end{bmatrix} \begin{bmatrix} \alpha \\ \beta \end{bmatrix} \notag \\ & \quad = \begin{bmatrix} \cos(\theta_0) & \sin(\theta_0) \\[1mm] \cos(\theta_R) \theta (z,R) - \sin(\theta_R) \theta' (z,R) & \cos(\theta_R) \phi (z,R) - \sin(\theta_R) \phi' (z,R) \end{bmatrix} \begin{bmatrix} \alpha \\[1mm] \beta \end{bmatrix} \notag \\ & \quad = {\mathcal U}(z,R,\theta_0,\theta_R) \begin{bmatrix} \alpha \\ \beta \end{bmatrix}, \label{2.40} \end{align} with ${\mathcal U}$ introduced in \eqref{2.2ca}. Solving \eqref{2.40} for $\alpha, \beta$ (and observing that $\det({\mathcal U}) = \Delta$ according to \eqref{2.2det}) then yields \eqref{2.37} and hence the special cases \eqref{2.38} and \eqref{2.39}. \end{proof} \section{Resolvent Formulas} \label{s3} In the principal result of this section, we will derive a formula for $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}}$ in terms of the resolvent of $\ensuremath{H_{\te_0,\te_R}}$ and the boundary traces $\gamma_{\theta'_0,\theta'_R}$. But first we focus on the special boundary data map given by \begin{equation} \ensuremath{\La_{\te_0,\te_R}}(z) = \Lateq (z), \quad \theta_0,\theta_R\in S_{2 \pi}, \quad z\in{\mathbb{C}}\backslash \sigma(\ensuremath{H_{\te_0,\te_R}}), \label{3.0} \end{equation} a generalization of the Dirichlet-to-Neumann map $\Lambda_{D,N}(z)=\ensuremath{\La_{0,0}}(z)$. Using the basis for solutions of \eqref{2.3} given in \eqref{2.5} by $u_{+,\theta_R}(z,\cdot), u_{-,\theta_0}(z,\cdot)$, and with $\theta_0^{\prime}=(\theta_0+\pi/2) \, \text{\rm mod} (2 \pi)$, $\theta_R^{\prime}=(\theta_R+\pi/2) \, \text{\rm mod} (2 \pi)$, equation \eqref{2.20} becomes \begin{align} & \ensuremath{\La_{\te_0,\te_R}} (z) = \big[\big(\ensuremath{\La_{\te_0,\te_R}} (z)\big)_{j,k}\big]_{1 \leq j,k \leq 2}, \quad z\in{\mathbb{C}}\backslash\sigma(\ensuremath{H_{\te_0,\te_R}}), \label{3.1} \\[1mm] \begin{split} & \big(\ensuremath{\La_{\te_0,\te_R}} (z)\big)_{1,1} = \frac{-\sin(\theta_0) + \cos(\theta_0)u_{+,\theta_R}'(z,0)}{\cos(\theta_0) + \sin(\theta_0)u_{+,\theta_R}'(z,0)}, \\ & \big(\ensuremath{\La_{\te_0,\te_R}} (z)\big)_{1,2} = \frac{-\sin(\theta_0)u_{-,\theta_0}(z,0) + \cos(\theta_0)u_{-,\theta_0}'(z,0)}{\cos(\theta_R) - \sin(\theta_R)u_{-,\theta_0}'(z,R)}, \\ & \big(\ensuremath{\La_{\te_0,\te_R}} (z)\big)_{2,1} = \frac{-\sin(\theta_R)u_{+,\theta_R}(z,R) - \cos(\theta_R)u_{+,\theta_R}'(z,R)}{\cos(\theta_0) + \sin(\theta_0)u_{+,\theta_R}'(z,0)}, \label{3.1a} \\ & \big(\ensuremath{\La_{\te_0,\te_R}} (z)\big)_{2,2} = \frac{-\sin(\theta_R) - \cos(\theta_R)u_{-,\theta_0}'(z,R)}{\cos(\theta_R) - \sin(\theta_R)u_{-,\theta_0}'(z,R)}. \end{split} \end{align} Next, consider \begin{equation}\label{3.2} \alpha(\xi)=\begin{bmatrix}\cos(\xi) & \sin(\xi)\end{bmatrix},\quad \xi\in S_{2 \pi}, \end{equation} and let $\Psi(z,\cdot \, ;x_0,\alpha(\xi))$ denote the fundamental matrix of solutions of \eqref{2.3} given by \begin{align}\label{3.3} \Psi(z,\cdot \, ;x_0,\alpha(\xi))&=\begin{bmatrix}\Theta(z,\cdot \, ;x_0,\alpha(\xi)) & \Phi(z,\cdot \, ;x_0,\alpha(\xi)) \end{bmatrix} \notag \\ &=\begin{bmatrix}\vartheta(z,\cdot \, ;x_0,\alpha(\xi))&\varphi(z,\cdot \, ;x_0,\alpha(\xi))\\ \vartheta'(z,\cdot \, ;x_0\alpha(\xi))&\varphi'(z,\cdot \, ;x_0,\alpha(\xi))\end{bmatrix}, \\ \intertext{in particular,} \Psi(z,x_0;x_0,\alpha(\xi))&=\left[\negthickspace\begin{array}{rr}\cos(\xi)&-\sin(\xi)\\ \sin(\xi)&\cos(\xi) \end{array}\negthickspace\right],\quad x_0\in[0,R]. \notag \end{align} In addition, set \begin{equation} \beta(\eta)=\begin{bmatrix}\cos(\eta) & -\sin(\eta)\end{bmatrix} = \alpha(-\eta), \quad \eta\in S_{2 \pi}. \label{3.3a} \end{equation} As shown in \cite[Section 2]{CG02} and \cite{CG06}, one can then introduce for all $x_0,y_0\in[0,R]$, $x_0\ne y_0$, the \emph{Weyl--Titchmarsh function}, $m(z;x_0,\alpha(\xi);y_0,\beta(\eta))$, given by \begin{align}\label{3.4} & m(z;x_0,\alpha(\xi);y_0,\beta(\eta)) \notag \\ & \quad = -[\beta(\eta)\Phi(z,y_0;x_0,\alpha(\xi))]^{-1}[\beta(\eta)\Theta(z,y_0;x_0,\alpha(\xi))] \notag \\[1mm] & \quad =-\dfrac{\cos(\eta)\vartheta(z,y_0;x_0,\alpha(\xi)) - \sin(\eta)\vartheta'(z,y_0;x_0,\alpha(\xi))}{\cos(\eta)\varphi(z,y_0;x_0,\alpha(\xi)) - \sin(\eta)\varphi'(z,y_0;x_0,\alpha(\xi))},\\[1mm] &\hspace*{4.6cm} \xi,\eta\in S_{2 \pi}, \; z\in{\mathbb{C}}\backslash\sigma(H_{\xi,\eta}). \notag \end{align} Then, with $\psi(z,\cdot \, ;x_0,\alpha(\xi);y_0,\beta(\eta))$ defined by \begin{align}\label{3.5} \begin{split} & \psi(z,\cdot \, ;x_0,\alpha(\xi);y_0,\beta(\eta)) \\ & \quad = \vartheta(z,\cdot \, ;x_0,\alpha(\xi))+ \varphi(z,\cdot \, ;x_0,\alpha(\xi))m(z;x_0,\alpha(\xi);y_0,\beta(\eta)), \end{split} \end{align} $\varphi(z,\cdot \, ;x_0,\alpha(\xi))$ and $\psi(z,\cdot \, ;x_0,\alpha(\xi);y_0,\beta(\eta))$ are linearly independent solutions of \eqref{2.3} satisfying \begin{align} \cos(\xi)\varphi(z,x_0;x_0,\alpha(\xi)) + \sin(\xi)\varphi'(z,x_0;x_0,\alpha(\xi))&=0,\label{3.6}\\ \cos(\eta)\psi(z,y_0;x_0,\alpha(\xi);y_0,\beta(\eta)) - \sin(\eta)\psi'(z,y_0;x_0,\alpha(\xi);y_0,\beta(\eta))&=0. \label{3.7} \end{align} The following result has been proved in \cite[Lemma 2.10]{CG02} in a self-adjoint context, but self-adjointness is of no relevance for the result \eqref{3.8} below. For the convenience of the reader we reproduce its proof here. \begin{lemma} [\cite{CG02}, Lemma 2.10] \label{l3.1} Suppose $\xi, \eta \in S_{2 \pi}$, $z\in{\mathbb{C}}\big\backslash\big(\sigma(H_{\xi,\eta}) \cup \sigma(H_{0,\eta})\big)$, let $\alpha(\xi)$ and $\beta(\eta)$ be defined as in \eqref{3.2} and \eqref{3.3a}, and suppose that $x_0, y_0 \in [0,R]$. Then, the following linear fractional transformation holds, \begin{equation}\label{3.8} m(z;x_0,\alpha(\xi);y_0,\beta(\eta))=\dfrac{-\sin(\xi) + \cos(\xi)m(z;x_0,\alpha(0);y_0,\beta(\eta))} {\cos(\xi) + \sin(\xi)m(z;x_0,\alpha(0);y_0,\beta(\eta))}, \end{equation} with $m$ defined as in \eqref{3.4}. \end{lemma} \begin{proof} With $\psi(z,\cdot \, ;x_0,\alpha(\xi);y_0,\beta(\eta))$ defined in \eqref{3.5}, by \eqref{3.7} one obtains \begin{equation \beta(\eta)\begin{bmatrix}\psi(z,y_0;x_0,\alpha(0);y_0,\beta(\eta)) \\ \psi'(z,y_0;x_0,\alpha(0);y_0,\beta(\eta)) \end{bmatrix}= \beta(\eta)\begin{bmatrix}\psi(z,y_0;x_0,\alpha(\xi);y_0,\beta(\eta)) \\ \psi'(z,y_0;x_0,\alpha(\xi);y_0,\beta(\eta)) \end{bmatrix}=0, \end{equation} and as a consequence, for some $c(z)\in {\mathbb{C}}\backslash\{0\}$, \begin{equation \begin{bmatrix}\psi(z,y_0;x_0,\alpha(\xi);y_0,\beta(\eta)) \\ \psi'(z,y_0;x_0,\alpha(\xi);y_0,\beta(\eta)) \end{bmatrix}= c(z)\begin{bmatrix}\psi(z,y_0;x_0,\alpha(0);y_0,\beta(\eta)) \\ \psi'(z,y_0;x_0,\alpha(0);y_0,\beta(\eta)) \end{bmatrix}. \end{equation} By the uniqueness of solutions for the initial value problem associated with \eqref{2.3}, \begin{equation}\label{3.13} \begin{bmatrix}\psi(z,x;x_0,\alpha(\xi);y_0,\beta(\eta)) \\ \psi'(z,x;x_0,\alpha(\xi);y_0,\beta(\eta)) \end{bmatrix}= c(z)\begin{bmatrix}\psi(z,x;x_0,\alpha(0);y_0,\beta(\eta)) \\ \psi'(z,x;x_0,\alpha(0);y_0,\beta(\eta)) \end{bmatrix}, \quad x\in{[0,R]}. \end{equation} Next, one notes that \begin{align}\label{3.14} &\begin{bmatrix}\psi(z,x_0;x_0,\alpha(\xi);y_0,\beta(\eta)) \\ \psi'(z,x_0;x_0,\alpha(\xi);y_0,\beta(\eta)) \end{bmatrix}= \Psi(z,x_0;x_0,\alpha(\xi))\begin{bmatrix}1\\ m(z;x_0,\alpha(\xi);y_0,\beta(\eta)) \end{bmatrix} \notag \\ & \quad =\begin{bmatrix}\cos(\xi)& -\sin(\xi)\\ \sin(\xi)& \cos(\xi) \end{bmatrix} \begin{bmatrix}1\\ m(z;x_0,\alpha(\xi);y_0,\beta(\eta)) \end{bmatrix}, \end{align} and similarly that \begin{align}\label{3.15} \begin{bmatrix}\psi(z,x_0;x_0,\alpha(0);y_0,\beta(\eta)) \\ \psi'(z,x_0;x_0,\alpha(0);y_0,\beta(\eta)) \end{bmatrix}&= \Psi(z,x_0;x_0,\alpha(0))\begin{bmatrix}1\\ m(z;x_0,\alpha(0);y_0,\beta(\eta)) \end{bmatrix} \notag \\ &=\begin{bmatrix}1\\ m(z;x_0,\alpha(0);y_0,\beta(\eta)) \end{bmatrix}. \end{align} By \eqref{3.13}--\eqref{3.15}, one concludes that \begin{equation \begin{bmatrix}1\\ m(z;x_0,\alpha(\xi);y_0,\beta(\eta)) \end{bmatrix}= c(z)\begin{bmatrix}\cos(\xi)& \sin(\xi)\\ -\sin(\xi)& \cos(\xi) \end{bmatrix} \begin{bmatrix}1\\ m(z;x_0,\alpha(0);y_0,\beta(\eta)) \end{bmatrix} \end{equation} implying \eqref{3.8}. \end{proof} With $\alpha(\xi)$ and $\beta(\eta)$ defined in \eqref{3.2} and \eqref{3.3a}, and with $m(z;x_0,\alpha(\xi);y_0,\beta(\eta))$ defined in \eqref{3.4}, for the next result we let \begin{align} &\alpha_0=\alpha(\theta_0)=\begin{bmatrix}\cos(\theta_0) & \sin(\theta_0) \end{bmatrix},\quad \beta_R=\beta(\theta_R)=\begin{bmatrix}\cos(\theta_R) & -\sin(\theta_R)\end{bmatrix},\label{3.9}\\ &m_{+,\theta_0}(z,\theta_R)=m(z;0,\alpha_0;R,\beta_R),\quad m_{-,\theta_R}(z,\theta_0)=m(z;R,\beta_R; 0,\alpha_0).\label{3.10} \end{align} \begin{theorem}\label{t3.2} Let $\theta_0, \theta_R \in S_{2 \pi}$ and $z\in{\mathbb{C}}\backslash\sigma(\ensuremath{H_{\te_0,\te_R}})$. Then, \begin{equation} \ensuremath{\La_{\te_0,\te_R}} (z) = \begin{bmatrix} m_{+,\theta_0}(z,\theta_R) & \ensuremath{\La_{\te_0,\te_R}} (z)_{1,2} \\[1mm] \ensuremath{\La_{\te_0,\te_R}} (z)_{2,1} & -m_{-,\theta_R}(z,\theta_0) \end{bmatrix}, \label{3.11} \end{equation} where \begin{align} &\ensuremath{\La_{\te_0,\te_R}} (z)_{1,2} = \ensuremath{\La_{\te_0,\te_R}} (z)_{2,1} \notag \\[1mm] \begin{split} & \quad = \dfrac{-\sin(\theta_0)u_{-,\theta_0}(z,0) + \cos(\theta_0)u_{-,\theta_0}'(z,0)}{\cos(\theta_R) - \sin(\theta_R)u_{-,\theta_0}'(z,R)} \\[1mm] & \quad = \dfrac{-\sin(\theta_R)u_{+,\theta_R}(z,R) - \cos(\theta_R)u_{+,\theta_R}'(z,R)}{\cos(\theta_0) + \sin(\theta_0)u_{+,\theta_R}'(z,0)}. \label{3.11a} \end{split} \end{align} \end{theorem} \begin{proof} We temporarily assume in addition that $z\notin \big(\sigma(H_{\theta_0,0}) \, \cup \, \sigma(H_{0,\theta_R})\big)$. If $x_0=0$, $y_0=R$, then with $\varphi(z,\cdot \, ;0,\alpha_0)$ defined in \eqref{3.3}, and with $\psi(z,\cdot \, ;0,\alpha_0; R,\beta_R)$ defined in \eqref{3.5}, one notes, by \eqref{3.6} and \eqref{3.7}, that \begin{align \varphi(z,\cdot \, ;0,\alpha_0) &= C_-(z) u_{-,\theta_0}(z,\cdot), \\ \psi(z,\cdot \, ;0,\alpha_0; R,\beta_R) &= C_+(z) u_{+,\theta_R}(z,\cdot), \end{align} for some $C_{\pm}(z)\in{\mathbb{C}}\backslash\{0\}$, where $u_{+,\theta_R}(z,\cdot), u_{-,\theta_0}(z,\cdot)$ represents the basis for the solutions of \eqref{2.3} described in \eqref{2.5}. With $m_{+,\theta_0}(z,\theta_R)$ defined in \eqref{3.10}, one notes when $\theta_0=0$ that \begin{equation m_{+,0}(z,\theta_R) =\dfrac{\psi'(z,0;0,\alpha_0; R,\beta_R)}{\psi(z,0;0,\alpha_0; R,\beta_R)}\bigg|_{\theta_0=0} = u_{+,\theta_R}'(z,0); \label{3.11b} \end{equation} hence by \eqref{3.8} that \begin{align} \begin{split} m_{+,\theta_0}(z,\theta_R) &= \dfrac{-\sin(\theta_0) + \cos(\theta_0)m_{+,0}(z,\theta_R)} {\cos(\theta_0) + \sin(\theta_0)m_{+,0}(z,\theta_R)} \\[1mm] &= \dfrac{-\sin(\theta_0) + \cos(\theta_0)u_{+,\theta_R}'(z,0)}{\cos(\theta_0) + \sin(\theta_0)u_{+,\theta_R}'(z,0)}. \label{3.24} \end{split} \end{align} By Theorem \ref{t2.3}, $\ensuremath{\La_{\te_0,\te_R}}$ is invariant with respect to a change of basis for \eqref{2.3}; thus the $(1,1)$-entry of $\ensuremath{\La_{\te_0,\te_R}}$, provided in \eqref{3.1a}, equals $m_{+,\theta_0}(z,\theta_R)$. Next, if $x_0=R$, $y_0=0$, then with $\varphi(z,\cdot \, ;R,\beta_R)$ defined in \eqref{3.3}, and with $\psi(z,\cdot \, ;R,\beta_R;0,\alpha_0)$ defined in \eqref{3.5}, one notes, by \eqref{3.6} and \eqref{3.7}, that \begin{align \varphi(z,\cdot \, ;R,\beta_R) &= D_+(z) u_{+,\theta_R}(z,\cdot), \\ \psi(z,\cdot \, ;R,\beta_R;0,\alpha_0) &= D_-(z) u_{-,\theta_0}(z,\cdot), \end{align} for some $D_{\pm}(z)\in{\mathbb{C}}\backslash\{0\}$, where $u_{+,\theta_R}(z,\cdot), u_{-,\theta_0}(z,\cdot)$ again denotes the basis for the solutions of \eqref{2.3} described in \eqref{2.5}. With $m_{-,\theta_R}(z,\theta_0)$ defined in \eqref{3.10}, one now obtains when $\theta_R=0$ that, \begin{equation m_{-,0}(z,\theta_0) =\dfrac{\psi'(z,R;R,\beta_R;0,\alpha_0)}{\psi(z,R;R,\beta_R;0,\alpha_0)}\bigg|_{\theta_R=0} = u_{-,\theta_0}'(z,R); \label{3.25} \end{equation} hence by \eqref{3.8} that \begin{align} \begin{split} m_{-,\theta_R}(z,\theta_0) &= \dfrac{\sin(\theta_R) + \cos(\theta_R)m_{-,0}(z,\theta_0)} {\cos(\theta_R) - \sin(\theta_R)m_{-,0}(z,\theta_0)} \\[1mm] &= \dfrac{\sin(\theta_R) + \cos(\theta_R)u_{-,\theta_0}'(z,R)}{\cos(\theta_R) - \sin(\theta_R)u_{-,\theta_0}'(z,R)}. \label{3.28} \end{split} \end{align} Again, by the invariance of $\ensuremath{\La_{\te_0,\te_R}}$ with respect to a change of basis for solutions of \eqref{2.3}, the $(2,2)$-entry of $\ensuremath{\La_{\te_0,\te_R}}$, provided in \eqref{3.1a}, is given by $-m_{-,\theta_R}(z,\theta_0)$. To see that the off-diagonal elements of \eqref{3.11} are equal, we introduce \begin{equation}\label{3.17} W(z) = W(u_{+,\theta_R}(z,\cdot), u_{-,\theta_0}(z,\cdot)), \quad \theta_0,\theta_R \in S_{2 \pi}, \; z\in{\mathbb{C}}\backslash\sigma(\ensuremath{H_{\te_0,\te_R}}), \end{equation} where $u_{+,\theta_R}(z,\cdot), u_{-,\theta_0}(z,\cdot)$, is the basis for the solutions of \eqref{2.3} as described in \eqref{2.5}. A straightforward computation, then yields \begin{align} & [-\sin(\theta_0)u_{-,\theta_0}(z,0) + \cos(\theta_0)u_{-,\theta_0}'(z,0)] [\cos(\theta_0) + \sin(\theta_0)u_{+,\theta_R}'(z,0)] \notag \\ & \quad = W(z) \label{3.18} \\ & \quad = [\cos(\theta_R) - \sin(\theta_R)u_{-,\theta_0}'(z,R)] [-\sin(\theta_R)u_{+,\theta_R}(z,R) - \cos(\theta_R)u_{+,\theta_R}'(z,R)]. \notag \end{align} Indeed, the first equality in \eqref{3.18} follows by inserting the second term in \eqref{2.5a} into \eqref{2.5c}; similarly, the second equality in \eqref{3.18} follows by inserting the second term in \eqref{2.5b} into \eqref{2.5d}. Equation \eqref{3.18} immediately yields \eqref{3.11a}. Finally, a meromorphic continuation with respect to $z$ then removes the additional assumption $z\notin \big(\sigma(H_{\theta_0,0})\cup \sigma(H_{0,\theta_R})\big)$, completing the proof. \end{proof} \begin{remark} \label{r3.3} Assume the special self-adjoint case where $V$ is real-valued and $\theta_0, \theta_R \in [0, 2\pi)$. By Marchenko's fundamental uniqueness result \cite[Ch.\ 2]{Ma73}, one observes the inverse spectral theory fact that each of the two diagonal terms of $\ensuremath{\La_{\te_0,\te_R}} (\cdot)$ already uniquely determines the potential coefficient $V(x)$ for a.e.\ $x\in [0,R]$. In addition, the known leading asymptotic behavior of $m_{+,\theta_0}(z,\theta_R)$ as $z \to i \, \infty$, \begin{align} \begin{split} m_{+,\theta_0}(z,\theta_R)& \underset{z \to i \infty}{\longrightarrow} \cot(\theta_0) + o(1), \quad \theta_0 \in [0, 2\pi)\backslash\{0,\pi\}, \\ m_{+,0}(z,\theta_R) & \underset{z \to i \infty}{\longrightarrow} i z^{1/2} + o\big(z^{1/2}\big), \quad \theta_0 \in \{0,\pi\}, \end{split} \end{align} determines the boundary condition parameter $\theta_0$, and similarly, that of $m_{-,\theta_R}(z,\theta_0)$ determines $\theta_R$. These facts are obviously not shared by the usual $2 \times 2$ matrix-valued Weyl--Titchmarsh function to be briefly discussed in our final Section \ref{s7} and hence demonstrates a stark contrast between these two matrix-valued functions, even though, both are Herglotz functions. The Herglotz property is well-known to be an intrinsic property of the $2 \times 2$ matrix-valued Weyl--Titchmarsh function (cf., e.g., \cite[Sect.\ 9.5]{CL85}), \cite[Sect.\ II.8]{LS75}, \cite[Sect.\ 6.5]{Pe88}, \cite[Ch.\ III]{Ti62}) and will also be verified for $\ensuremath{\La_{\te_0,\te_R}} (\cdot)$ as a special case of the principal result in Theorem \ref{t4.6} (cf.\ \eqref{4.76}). \end{remark} Next we turn to the asymptotic behavior of $\ensuremath{\La_{\te_0,\te_R}} (z)$ as $|z|\to\infty$, $\Im(z^{1/2})>0$: \begin{lemma} \label{l3.4} Suppose that $V\in L^1((0,R); dx)$ and let $\theta_0, \theta_R\in S_{2 \pi}$. Then, \begin{align} & \ensuremath{\La_{\te_0,\te_R}} (z) \underset{\substack{|z|\to\infty\\ \Im(z^{1/2})>0}}{=} \begin{bmatrix} \cot(\theta_0) & \frac{2i z^{-1/2} e^{i z^{1/2} R}}{\sin(\theta_0)\sin(\theta_R)} \\ \frac{2i z^{-1/2} e^{i z^{1/2} R}}{\sin(\theta_0)\sin(\theta_R)} & - \cot(\theta_0) \end{bmatrix} \notag \\[1mm] & \quad + \begin{bmatrix} O(|z|^{-1/2} ) & O\big(|z|^{-1} e^{-\Im(z^{1/2}) R}\big) \\ O\big(|z|^{-1} e^{-\Im(z^{1/2}) R}\big) & O(|z|^{-1/2}) \end{bmatrix}, \quad \theta_0 \neq 0, \, \theta_R \neq 0, \\ & \Lambda_{0,\theta_R} (z) \underset{\substack{|z|\to\infty\\ \Im(z^{1/2})>0}}{=} \begin{bmatrix} i z^{1/2} & - \frac{2 e^{i z^{1/2} R}}{\sin(\theta_R)} \\ - \frac{2 e^{i z^{1/2} R}}{\sin(\theta_R)} & - \cot(\theta_R) \end{bmatrix} \notag \\[1mm] & \quad + \begin{bmatrix} O(1) & O\big(|z|^{-1/2} e^{-\Im(z^{1/2}) R}\big) \\ O\big(|z|^{-1/2} e^{-\Im(z^{1/2}) R}\big) & O(|z|^{-1/2}) \end{bmatrix}, \quad \theta_0 = 0, \, \theta_R \neq 0, \\ & \Lambda_{\theta_0,0} (z) \underset{\substack{|z|\to\infty\\ \Im(z^{1/2})>0}}{=} \begin{bmatrix} \cot(\theta_0) & - \frac{2 e^{i z^{1/2} R}}{\sin(\theta_0)} \\ - \frac{2 e^{i z^{1/2} R}}{\sin(\theta_0)} & - i z^{1/2} \end{bmatrix} \notag \\[1mm] & \quad + \begin{bmatrix} O(|z|^{-1/2}) & O\big(|z|^{-1/2} e^{-\Im(z^{1/2}) R}\big) \\ O\big(|z|^{-1/2} e^{-\Im(z^{1/2}) R}\big) & O(1) \end{bmatrix}, \quad \theta_0 \neq 0, \, \theta_R = 0, \\ & \Lambda_{0,0} (z) \underset{\substack{|z|\to\infty\\ \Im(z^{1/2})>0}}{=} \begin{bmatrix} i z^{1/2} & - 2i z^{1/2} e^{i z^{1/2} R} \\ - 2i z^{1/2} e^{i z^{1/2} R} & - i z^{1/2} \end{bmatrix} \notag \\[1mm] & \quad + \begin{bmatrix} O(1) & O\big(e^{-\Im(z^{1/2}) R}\big) \\ O\big(e^{-\Im(z^{1/2}) R}\big) & O(1) \end{bmatrix}, \quad \theta_0 = 0, \, \theta_R = 0. \end{align} \end{lemma} \begin{proof} This follows from an elementary computation upon inserting \eqref{2.38} (resp., \eqref{2.39}) into \eqref{3.24} (resp., \eqref{3.28}) and using the asymptotic expansions \eqref{2.2e}. \end{proof} For the principal result of this section, an explicit formula for $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}}(z)$ in terms of the resolvent $(\ensuremath{H_{\te_0,\te_R}}- z I \big)^{-1}$ of $\ensuremath{H_{\te_0,\te_R}}$ and the boundary traces $\gamma_{\theta'_0,\theta'_R}$, we recall the Green's function associated with the operator $\ensuremath{H_{\te_0,\te_R}}$ in \eqref{2.2a}, \begin{align} & \ensuremath{G_{\te_0,\te_R}}(z,x,x') = (\ensuremath{H_{\te_0,\te_R}} -z I)^{-1}(x,x') \notag \\ & \quad = \frac{1}{W(u_{+,\theta_R}(z,\cdot), u_{-,\theta_0}(z,\cdot))} \begin{cases} u_{-,\theta_0}(z,x')u_{+,\theta_R}(z,x), & 0\le x'\le x,\\[1mm] u_{-,\theta_0}(z,x)u_{+,\theta_R}(z,x'), & 0\le x\le x', \end{cases} \label{3.32} \\[1mm] &\hspace*{5.9cm} z\in{\mathbb{C}}\backslash\sigma(\ensuremath{H_{\te_0,\te_R}}), \; x, x' \in {[0,R]}. \notag \end{align} Here $u_{+,\theta_R}(z,\cdot), u_{-,\theta_0}(z,\cdot)$ is a basis for solutions of \eqref{2.3} as described in \eqref{2.5} and $I = I_{L^2((0,R); dx)}$ abbreviates the identity operator in $L^2((0,R); dx)$. Thus, one obtains \begin{align} \begin{split} \big((\ensuremath{H_{\te_0,\te_R}}- z I)^{-1}g\big)(x) = \int_0^R \, dx'\, \ensuremath{G_{\te_0,\te_R}}(z,x,x')g(x'),& \\ g\in L^2((0,R); dx), \; z\in{\mathbb{C}}\backslash\sigma(\ensuremath{H_{\te_0,\te_R}}), \; x\in(0,R).& \label{3.33} \end{split} \end{align} Equation \eqref{3.32} is a crucial input in the proof of the following fundamental result: \begin{theorem} \label{t3.5} Assume that $\theta_0, \theta_R, \theta_0', \theta_R' \in S_{2 \pi}$, let $\ensuremath{H_{\te_0,\te_R}}$ be defined as in \eqref{2.2a}, suppose that $z\in{\mathbb{C}}\backslash\sigma(\ensuremath{H_{\te_0,\te_R}})$, and let $S_{\theta_0' -\theta_0,\theta_R' -\theta_R}$ be defined according to \eqref{2.14}. Then \begin{equation} \ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z) S_{\theta_0' -\theta_0,\theta_R'-\theta_R} =\gamma_{\theta_0',\theta_R'} \big[\gamma_{\overline{\theta_0'},\overline{\theta_R'}} ((\ensuremath{H_{\te_0,\te_R}})^* - {\overline z} I)^{-1}\big]^*. \label{3.33aa} \end{equation} In particular, with $\theta_0^{\prime}=(\theta_0+\pi/2) \, \text{\rm mod} (2 \pi)$, $\theta_R^{\prime}=(\theta_R+\pi/2) \, \text{\rm mod} (2 \pi)$, one obtains \begin{equation} \ensuremath{\La_{\te_0,\te_R}} (z) =\ensuremath{\widehat \ga_{\te_0,\te_R}} \big[\widehat \gamma_{\overline{\theta_0},\overline{\theta_R}}((\ensuremath{H_{\te_0,\te_R}})^* - {\overline z} I)^{-1}\big]^*, \label{3.33a} \end{equation} where \begin{equation} \ensuremath{\widehat \ga_{\te_0,\te_R}}=\gateq. \end{equation} As a consequence, \begin{equation} \Lambda_{(\theta_0 + \pi) \, \text{\rm mod} (2\pi), (\theta_R + \pi) \, \text{\rm mod} (2\pi)} (z) = \ensuremath{\La_{\te_0,\te_R}} (z), \quad \theta_0, \theta_R\in S_{2 \pi}. \label{3.33A} \end{equation} \end{theorem} \begin{proof} We start by noting that the Green's function $G^*_{\theta_0,\theta_R}(z,x,x')$ of $(\ensuremath{H_{\te_0,\te_R}})^*$ is given by \begin{align}\label{3.32a} \begin{split} \ensuremath{G_{\te_0,\te_R}}^*(z,x,x') &= \big((\ensuremath{H_{\te_0,\te_R}})^* -z I \big)^{-1}(x,x') \notag \\ & = \frac{1}{W^*(z)} \begin{cases} \overline{u_{-,\theta_0}(\overline z,x')} \, \overline{u_{+,\theta_R}(\overline z,x)}, & 0\le x'\le x,\\[2mm] \overline{u_{-,\theta_0}(\overline z,x)}\, \overline{u_{+,\theta_R}(\overline z,x')}, & 0\le x\le x', \end{cases} \\[1mm] &\hspace*{2.25cm} z\in{\mathbb{C}}\backslash\sigma\big((\ensuremath{H_{\te_0,\te_R}})^*\big), \; x, x' \in {[0,R]}, \end{split} \end{align} where $u_{+,\theta_R}(z,\cdot), u_{-,\theta_0}(z,\cdot)$ is a basis for solutions of \eqref{2.3} as described in \eqref{2.5}, and $W^*(z)$ is defined by \begin{equation}\label{3.17a} W^*(z) = W\big(\overline{u_{+,\theta_R}(\overline z,\cdot)}, \overline{u_{-,\theta_0}(\overline z,\cdot)}\big), \quad z\in{\mathbb{C}}\backslash\sigma(\ensuremath{H_{\te_0,\te_R}}). \end{equation} In particular, we note the fact that \begin{equation}\label{3.38} \overline{W^*(\overline z)} = W(z), \end{equation} with $W(z)$ the Wronskian defined in \eqref{3.17}. Thus, one obtains \begin{align} \begin{split} \big(((\ensuremath{H_{\te_0,\te_R}})^*- z I)^{-1}g\big)(x) = \int_0^R \, dx'\, \ensuremath{G_{\te_0,\te_R}}^*(z,x,x') g(x'),& \\ g\in L^2((0,R); dx), \; z\in{\mathbb{C}}\backslash\sigma(\ensuremath{H_{\te_0,\te_R}}), \; x\in(0,R).& \label{3.38a} \end{split} \end{align} Next, for $g\in L^2((0,R); dx)$ let \begin{align \ensuremath{\widehat g}(\overline z,x) & = \big(((\ensuremath{H_{\te_0,\te_R}})^*- {\overline z} I)^{-1} g\big)(x) \notag \\ & = \frac{1}{\overline{W(z)}} \bigg(\overline{u_{+,\theta_R}(z,x)} \int_0^x\, dx'\, \overline{u_{-,\theta_0}(z,x')} g(x') \\ & \hspace*{1.75cm} + \overline{u_{-,\theta_0}(z,x)} \int_x^R\, dx'\, \overline{u_{+,\theta_R}(z,x')} g(x')\bigg). \notag \\ \intertext{Then one notes that} \begin{split} \ensuremath{\widehat g}'(\overline z,x) & = \frac{1}{\overline{W(z)}} \bigg(\overline{u_{+,\theta_R}'(z,x)} \int_0^x\, dx'\, \overline{u_{-,\theta_0}(z,x')} g(x') \\ & \hspace*{1.75cm} + \overline{u_{-,\theta_0}'(z,x)} \int_x^R\, dx'\, \overline{u_{+,\theta_R}(z,x')} g(x')\bigg), \end{split} \end{align} and, as a consequence, that \begin{align} \ensuremath{\widehat g}(\overline z,0)&=\frac{1}{\overline{W(z)}} \overline{u_{-,\theta_0}(z,0)} \int_0^R \, dx'\, \overline{u_{+,\theta_R}(z,x')} g(x'), \label{3.36} \\ \ensuremath{\widehat g}'(\overline z,0)&=\frac{1}{\overline{W(z)}} \overline{u_{-,\theta_0}'(z,0)} \int_0^R \, dx'\, \overline{u_{+,\theta_R}(z,x')} g(x'), \label{3.36a} \\ \ensuremath{\widehat g}(\overline z,R)& =\frac{1}{\overline{W(z)}} \overline{u_{+,\theta_R}(z,R)} \int_0^R \, dx'\, \overline{u_{-,\theta_0}(z,x')} g(x'), \label{3.37} \\ \ensuremath{\widehat g}'(\overline z,R)&=\frac{1}{\overline{W(z)}} \overline{u_{+,\theta_R}'(z,R)} \int_0^R \, dx'\, \overline{u_{-,\theta_0}(z,x')} g(x'). \label{3.37a} \end{align} In turn, this permits one to compute \begin{align} \begin{split} \gamma_{\overline{\theta_0'},\overline{\theta_R'}} ((\ensuremath{H_{\te_0,\te_R}})^* - {\overline z} I)^{-1} g &= \gamma_{\overline{\theta_0'},\overline{\theta_R'}} \ensuremath{\widehat g} (\ensuremath{\bar z},\cdot) \\ & = \begin{bmatrix} \cos(\overline{\theta_0'}) \widehat g (\overline z, 0) + \sin(\overline{\theta_0'}) {\widehat g}^{\prime}(\overline z, 0) \\[1mm] \cos(\overline{\theta_R'}) \widehat g(\overline z, R) - \sin(\overline{\theta_R'}) {\widehat g}^{\prime}(\overline z, R) \end{bmatrix} \label{3.42}. \end{split} \end{align} Using \eqref{3.36}--\eqref{3.42} one infers, with $[a_0 \;\, a_R]^\top \in {\mathbb{C}}^2$ and $(\cdot,\cdot)_{{\mathbb{C}}^2}$ denoting the scalar product in ${\mathbb{C}}^2$, that \begin{align} &\big(\gamma_{\overline{\theta_0'},\overline{\theta_R'}} ((\ensuremath{H_{\te_0,\te_R}})^* - {\overline z} I)^{-1} g, [a_0 \;\, a_R]^\top \big)_{{\mathbb{C}}^2} \notag \\ & \quad = \left( \begin{bmatrix} \cos(\overline{\theta_0'}) \widehat g (\overline z, 0) + \sin(\overline{\theta_0'}) {\widehat g}^{\prime}(\overline z, 0) \\[1mm] \cos(\overline{\theta_R'}) \widehat g(\overline z, R) - \sin(\overline{\theta_R'}) {\widehat g}^{\prime}(\overline z, R) \end{bmatrix}, \begin{bmatrix} a_0 \\[1mm] a_R \end{bmatrix} \right)_{{\mathbb{C}}^2} \notag \\ & \quad =\frac{1}{W(z)} \int_0^R \, dx' \, \overline{g(x')} \big(\cos(\theta_0') a_0 u_{-,\theta_0}(z,0) u_{+,\theta_R}(z,x') \notag \\ & \hspace*{1.3cm} + \sin(\theta_0') a_0 u_{-,\theta_0}^{\prime}(z,0) u_{+,\theta_R}(z,x') +\cos(\theta_R') a_R u_{+,\theta_R}(z,R) u_{-,\theta_0}(z,x') \notag \\ & \hspace*{1.3cm} -\sin(\theta_R') a_R u_{+,\theta_R}^{\prime}(z,R) u_{-,\theta_0}(z,x') \big). \end{align} Hence one concludes that \begin{align} &\Big(\big[\gamma_{\overline{\theta_0'},\overline{\theta_R'}} ((\ensuremath{H_{\te_0,\te_R}})^*- {\overline z} I)^{-1}\big]^* [a_0 \;\, a_R]^\top \Big)(x) \notag \\ &\quad =\frac{1}{W(z)} \Big(\big[\cos(\theta_0') u_{-,\theta_0}(z,0) + \sin(\theta_0') u_{-,\theta_0}^{\prime}(z,0)\big] a_0 u_{+,\theta_R}(z,x) \notag \\ & \hspace*{1.95cm} + \big[\cos(\theta_R') u_{+,\theta_R}(z,R) - \sin(\theta_R') u_{+,\theta_R}^{\prime}(z,R)\big] a_R u_{-,\theta_0}(z,x) \Big). \label{3.43} \end{align} Consequently, writing \begin{align} \begin{split} & \gamma_{\theta_0',\theta_R'} \big[\gamma_{\overline{\theta_0'},\overline{\theta_R'}} ((\ensuremath{H_{\te_0,\te_R}})^*- {\overline z} I)^{-1}\big]^* [a_0 \;\, a_R]^\top \\ & \quad = \begin{bmatrix} \big(\gamma_{\theta_0',\theta_R'} \big[\gamma_{\overline{\theta_0'},\overline{\theta_R'}} ((\ensuremath{H_{\te_0,\te_R}})^*- {\overline z} I)^{-1}\big]^* [a_0 \;\, a_R]^\top\big)_1 \\[1mm] \big(\gamma_{\theta_0',\theta_R'} \big[\gamma_{\overline{\theta_0'},\overline{\theta_R'}} ((\ensuremath{H_{\te_0,\te_R}})^*- {\overline z} I)^{-1}\big]^* [a_0 \;\, a_R]^\top\big)_2 \end{bmatrix}, \end{split} \end{align} it follows that \begin{align} & \big(\gamma_{\theta_0',\theta_R'} \big[\gamma_{\overline{\theta_0'},\overline{\theta_R'}} ((\ensuremath{H_{\te_0,\te_R}})^*- {\overline z} I)^{-1}\big]^* [a_0 \;\, a_R]^\top \big)_1 \notag \\ & \quad = \frac{1}{W(z)} \Big(\cos(\theta_0')\big[\cos(\theta_0') u_{-,\theta_0}(z,0) + \sin(\theta_0') u'_{-,\theta_0}(z,0)\big] a_0 \notag \\ & \qquad + \cos(\theta_0')\big[\cos(\theta_R') u_{+,\theta_R}(z,R) - \sin(\theta_R') u'_{+,\theta_R}(z,R)\big] a_R u_{-,\theta_0}(z,0) \notag \\ & \qquad + \sin(\theta_0')\big[\cos(\theta_0') u_{-,\theta_0}(z,0) + \sin(\theta_0') u'_{-,\theta_0}(z,0)\big] a_0 u'_{+,\theta_R}(z,0) \notag \\ & \qquad + \sin(\theta_0')\big[\cos(\theta_R') u_{+,\theta_R}(z,R) - \sin(\theta_R') u'_{+,\theta_R}(z,R)\big] a_R u'_{-,\theta_0}(z,0)\Big) , \\ \intertext{and} & \big(\gamma_{\theta_0',\theta_R'} \big[\gamma_{\overline{\theta_0'},\overline{\theta_R'}} ((\ensuremath{H_{\te_0,\te_R}})^*- {\overline z} I)^{-1}\big]^* [a_0 \;\, a_R]^\top \big)_2 \notag \\ & \quad = \frac{1}{W(z)} \Big(\cos(\theta_R')\big[\cos(\theta_0') u_{-,\theta_0}(z,0) + \sin(\theta_0') u'_{-,\theta_0}(z,0)\big] a_0 u_{+,\theta_R}(z,R) \notag \\ & \qquad + \cos(\theta_R')\big[\cos(\theta_R') u_{+,\theta_R}(z,R) - \sin(\theta_R') u'_{+,\theta_R}(z,R)\big] a_R \notag \\ & \qquad - \sin(\theta_R')\big[\cos(\theta_0') u_{-,\theta_0}(z,0) + \sin(\theta_0') u'_{-,\theta_0}(z,0)\big] a_0 u'_{+,\theta_R}(z,R) \notag \\ & \qquad - \sin(\theta_R')\big[\cos(\theta_R') u_{+,\theta_R}(z,R) - \sin(\theta_R') u'_{+,\theta_R}(z,R)\big] a_R u'_{-,\theta_0}(z,R)\Big). \end{align} Hence, writing \begin{align} \begin{split} & \gamma_{\theta_0',\theta_R'} \big[\gamma_{\overline{\theta_0'},\overline{\theta_R'}} ((\ensuremath{H_{\te_0,\te_R}})^*- {\overline z} I)^{-1}\big]^* \\ & \quad = \Big[\big(\gamma_{\theta_0',\theta_R'} \big[\gamma_{\overline{\theta_0'},\overline{\theta_R'}} ((\ensuremath{H_{\te_0,\te_R}})^*- {\overline z} I)^{-1}\big]^*\big)_{j,k}\Big]_{j,k=1,2}, \end{split} \end{align} one obtains \begin{align} \begin{split} & \big(\gamma_{\theta_0',\theta_R'} \big[\gamma_{\overline{\theta_0'},\overline{\theta_R'}} ((\ensuremath{H_{\te_0,\te_R}})^*- {\overline z} I)^{-1}\big]^*\big)_{1,1} \\ & \quad = \frac{1}{W(z)} \big[\cos(\theta_0') u_{-,\theta_0}(z,0) + \sin(\theta_0') u'_{-,\theta_0}(z,0)\big] \\ & \qquad \qquad \quad \times \big[\cos(\theta'_0) + \sin(\theta'_0) u'_{+,\theta_R}(z,0)\big], \\ & \big(\gamma_{\theta_0',\theta_R'} \big[\gamma_{\overline{\theta_0'},\overline{\theta_R'}} ((\ensuremath{H_{\te_0,\te_R}})^*- {\overline z} I \big)^{-1}\big]^*\big)_{1,2} \\ & \quad = \frac{1}{W(z)} \big[\cos(\theta_R') u_{+,\theta_R}(z,R) - \sin(\theta_R') u'_{+,\theta_R}(z,R)\big] \\ & \qquad \qquad \quad \times \big[\cos(\theta'_0) u_{-,\theta_0}(z,0) + \sin(\theta'_0) u'_{-,\theta_0}(z,0)\big], \\ & \big(\gamma_{\theta_0',\theta_R'} \big[\gamma_{\overline{\theta_0'},\overline{\theta_R'}} ((\ensuremath{H_{\te_0,\te_R}})^*- {\overline z} I)^{-1}\big]^*\big)_{2,1} \\ & \quad = \frac{1}{W(z)} \big[\cos(\theta_0') u_{-,\theta_0}(z,0) + \sin(\theta_0') u'_{-,\theta_0}(z,0)\big] \\ & \qquad \qquad \quad \times \big[\cos(\theta_R') u_{+,\theta_R}(z,R) - \sin(\theta_R') u'_{+,\theta_R}(z,R)\big], \\ & \big(\gamma_{\theta_0',\theta_R'} \big[\gamma_{\overline{\theta_0'},\overline{\theta_R'}} ((\ensuremath{H_{\te_0,\te_R}})^*- {\overline z} I)^{-1}\big]^*\big)_{2,2} \\ & \quad = \frac{1}{W(z)} \big[\cos(\theta_R') u_{+,\theta_R}(z,R) - \sin(\theta_R') u'_{+,\theta_R}(z,R)\big] \\ & \qquad \qquad \quad \times \big[\cos(\theta'_R) - \sin(\theta'_R) u'_{-,\theta_0}(z,R)\big]. \label{3.47} \end{split} \end{align} Employing \eqref{2.5c} and \eqref{2.5d} one finally concludes from \eqref{2.20a} that the expressions in \eqref{3.47} are equivalent to the following ones \begin{align} \begin{split} & \big(\gamma_{\theta_0',\theta_R'} \big[\gamma_{\overline{\theta_0'},\overline{\theta_R'}} ((\ensuremath{H_{\te_0,\te_R}})^*- {\overline z} I)^{-1}\big]^*\big)_{1,1} = \sin(\theta_0'-\theta_0) \Big(\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z)\Big)_{1,1}, \\ & \big(\gamma_{\theta_0',\theta_R'} \big[\gamma_{\overline{\theta_0'},\overline{\theta_R'}} ((\ensuremath{H_{\te_0,\te_R}})^*- {\overline z} I)^{-1}\big]^*\big)_{1,2} = \sin(\theta_R'-\theta_R) \Big(\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z)\Big)_{1,2}, \\ & \big(\gamma_{\theta_0',\theta_R'} \big[\gamma_{\overline{\theta_0'},\overline{\theta_R'}} ((\ensuremath{H_{\te_0,\te_R}})^*- {\overline z} I)^{-1}\big]^*\big)_{2,1} = \sin(\theta_0'-\theta_0) \Big(\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z)\Big)_{2,1}, \\ & \big(\gamma_{\theta_0',\theta_R'} \big[\gamma_{\overline{\theta_0'},\overline{\theta_R'}} ((\ensuremath{H_{\te_0,\te_R}})^*- {\overline z} I)^{-1}\big]^*\big)_{2,2} = \sin(\theta_R'-\theta_R) \Big(\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z)\Big)_{2,2}, \label{3.50} \end{split} \end{align} proving \eqref{3.33aa}. Finally, equation \eqref{3.33A} is a consequence of \eqref{2.2A} and \eqref{3.33a}. \end{proof} \begin{remark} \label{r3.6} A formula of the type \eqref{3.33a} for Dirichlet-to-Neumann maps associated with multi-dimensional Schr\"odinger operators was published by Amrein and Pearson \cite{AP04} in 2004. It has recently been extended in various directions in \cite{GM08}, \cite{GM09}, \cite{GM10}, \cite{GMZ07}. Formula \eqref{3.33a} for the Dirichlet-to-Neumann map $\Lambda_{\pi/2,\pi/2}$ in the special case $V=0$ has also been derived by Posilicano \cite[Example 5.1]{Po08}. \end{remark} \begin{remark} \label{r3.7} While it is tempting to view $\ensuremath{\ga_{\te_0,\te_R}}$ as an unbounded but densely defined operator on $L^2((0,R); dx)$ whose domain contains the space $C_0^\infty((0,R))$, one should note, in this case, that its adjoint $\ensuremath{\ga_{\te_0,\te_R}}^*$ is not densely defined. Indeed, the adjoint $\ensuremath{\ga_{\te_0,\te_R}}^*$ of $\ensuremath{\ga_{\te_0,\te_R}}$ would have to be an unbounded operator from ${\mathbb{C}}^2$ to $L^2((0,R); dx)$ such that \begin{equation} (\ensuremath{\ga_{\te_0,\te_R}} f,g)_{{\mathbb{C}}^2} = (f,\ensuremath{\ga_{\te_0,\te_R}}^* g)_{L^2((0,R); dx)} \, \text{ for all }\, f\in\text{\rm{dom}}(\ensuremath{\ga_{\te_0,\te_R}}),\; g\in\text{\rm{dom}}(\ensuremath{\ga_{\te_0,\te_R}}^*). \label{3.55} \end{equation} In particular, choosing $f\in C_0^\infty((0,R))$, in which case $\ensuremath{\ga_{\te_0,\te_R}} f =0$, one concludes that \begin{equation} (f,\ensuremath{\ga_{\te_0,\te_R}}^* g)_{L^2((0,R); dx)}=0 \, \text{ for all } \, f\in C_0^\infty((0,R)). \end{equation} Thus, one obtains $\ensuremath{\ga_{\te_0,\te_R}}^* g = 0$ for all $g\in\text{\rm{dom}}(\ensuremath{\ga_{\te_0,\te_R}}^*)$. Since obviously $\ensuremath{\ga_{\te_0,\te_R}} \neq 0$, \eqref{3.55} implies $\text{\rm{dom}}(\ensuremath{\ga_{\te_0,\te_R}}^*)=\{0\}$ and hence $\ensuremath{\ga_{\te_0,\te_R}}$ is not a closable linear operator in $L^2((0,R); dx)$. This is the reason for our careful choice of notation in \eqref{3.33aa} and \eqref{3.33a}. \end{remark} We conclude this section by providing an explicit example in the special case $V=0$ a.e.\ on $[0,R]$. For notational purposes we add the superscript $(0)$ to the corresponding quantities below: \begin{example}\label{e3.8} Let $V=0$ a.e.\ on $[0,R]$, $x_0 \in (0,R)$, $\theta_0, \theta_R, \theta \in S_{2 \pi}$, $x, x', s \in [0,R]$, and let \begin{align} &f(z,s,\alpha,\beta)=f(z,s,\beta,\alpha) \\ &\quad =z\sin(\alpha)\sin(\beta)\sin(\sqrt{z}s) +\sqrt{z}\sin(\alpha+\beta)\cos(\sqrt{z}s) - \cos(\alpha)\cos(\beta)\sin(\sqrt{z}s), \notag \\ &g(z,s,\alpha,\beta)=f(z,s,\alpha+\pi/2,\beta) \\ &\quad = z\cos(\alpha)\sin(\beta)\sin(\sqrt{z}s) +\sqrt{z}\cos(\alpha+\beta)\cos(\sqrt{z}s) + \sin(\alpha)\cos(\beta)\sin(\sqrt{z}s). \notag \end{align} Then, \begin{align} &u^{(0)} (z,x,(\theta_0,c_0),(\theta_R,c_R)) = \frac{c_0f(z,R-x,0,\theta_R) + c_Rf(z,x,\theta_0,0)}{f(z,R,\theta_0,\theta_R)}, \\ &u_{+,\theta_R}^{(0)} (z,x) = u(z,\cdot \, ;(0,1),(\theta_R,0)) = \frac{f(z,R-x,0,\theta_R)}{f(z,R,0,\theta_R)}, \\ &u_{+,\theta_R}^{(0)\;'} (z,x)= \frac{f'(z,R-x,0,\theta_R)}{f(z,R,0,\theta_R)} = \frac{f(z,R-x,\pi/2,\theta_R)}{f(z,R,0,\theta_R)}, \\ &u_{-,\theta_0}^{(0)} (z,x) = u(z,\cdot \, ;(\theta_0,0),(0,1)) = \frac{f(z,x,\theta_0,0)}{f(z,R,\theta_0,0)}, \\ &u_{-,\theta_0}^{(0)\; '}(z,x) = \frac{f'(z,x,\theta_0,0)}{f(z,R,\theta_0,0)} = -\frac{f(z,x,\theta_0,\pi/2)}{f(z,R,\theta_0,0)}, \\ &W(u^{(0)}_{+,\theta_R}(z,\cdot), u^{(0)}_{-,\theta_0}(z,\cdot)) = \frac{-\sqrt{z}f(z,R,\theta_0,\theta_R)}{f(z,R,\theta_0,0)f(z,R,0,\theta_R)}, \\ & m_{+,\theta_0}^{(0)} (z,\theta_R) = \frac{g(z,R,\theta_0,\theta_R)}{f(z,R,\theta_0,\theta_R)}, \\ & m_{-,\theta_R}^{(0)} (z,\theta_0) = - \frac{g(z,R,\theta_R,\theta_0)}{f(z,R,\theta_R,\theta_0)}, \\ & \Lambda_{\theta_0,\theta_R}^{(0)} (z) = \begin{pmatrix} \frac{g(z,R,\theta_0,\theta_R)}{f(z,R,\theta_0,\theta_R)} & \frac{-\sqrt{z}}{f(z,R,\theta_0,\theta_R)} \\ \frac{-\sqrt{z}}{f(z,R,\theta_0,\theta_R)} & \frac{g(z,R,\theta_R,\theta_0)}{f(z,R,\theta_0,\theta_R)} \end{pmatrix}, \label{3.72} \notag \\ &m_{+,0}^{(0)} (z, x_0,\theta_R) = \frac{u_{+,\theta_R}^{(0)\;'} (z,x_0)}{u_{+,\theta_R}^{(0)} (z,x_0)} = \frac{g(z,R-x_0,0,\theta_R)}{f(z,R-x_0,0,\theta_R)}, \\ &m_{-,0}^{(0)} (z, x_0,\theta_0) = \frac{u_{-,\theta_0}^{(0)\; '}(z,x_0)}{u_{-,\theta_0}^{(0)}(z,x_0)} = -\frac{g(z,x_0,0,\theta_0)}{f(z,x_0,0,\theta_0)}, \\ &G_{\theta_0,\theta_R}^{(0)} (z,x,x') = \big(\ensuremath{H_{\te_0,\te_R}}^{(0)} -z I\big)^{-1}(x,x') \notag \\ &\quad = \frac{1}{W(u^{(0)}_{+,\theta_R}(z,\cdot), u^{(0)}_{-,\theta_0}(z,\cdot))} \begin{cases} u^{(0)}_{-,\theta_0}(z,x')u^{(0)}_{+,\theta_R}(z,x), & 0\le x'\le x,\\[1mm] u^{(0)}_{-,\theta_0}(z,x)u^{(0)}_{+,\theta_R}(z,x'), & 0\le x\le x', \end{cases} \notag \\[10pt] &\quad = \begin{cases} \dfrac{f(z,x',\theta_0,0)f(z,R-x,0,\theta_R)}{-\sqrt{z}f(z,R,\theta_0,\theta_R)},& 0\le x'\le x,\\[10pt] \dfrac{f(z,x,\theta_0,0)f(z,R-x',0,\theta_R)}{-\sqrt{z}f(z,R,\theta_0,\theta_R)},& 0\le x\le x', \end{cases} \quad z\in{\mathbb{C}}\backslash\sigma\big(\ensuremath{H_{\te_0,\te_R}}^{(0)}\big). \end{align} \end{example} The special case of the Neumann-to-Dirichlet map, $\Lambda^{(0)}_{\pi/2,\pi/2}$ in \eqref{3.72}, was computed in \cite[Example\ 4.1]{DM92}, and more recently, the special case of the Dirichlet-to-Neumann map, $\Lambda^{(0)}_{0,0}$ in \eqref{3.72}, was computed in \cite[Example\ 5.1]{Po08}. \section{Linear Fractional Transformations and the Herglotz Property in the Self-Adjoint Case} \label{s4} The principal purpose of this section is to prove that $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}}(z)$ and $\ensuremath{\La_{\de_0,\de_R}^{\de_0^{\prime},\de_R^{\prime}}}(z)$ satisfy a linear fractional transformation. As a consequence we will show that $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}}(z) S_{\theta_0'-\theta_0,\theta_R'-\theta_R}$ is a $2 \times 2$ matrix-valued Herglotz function in the special self-adjoint case where $V$ and $\theta_0, \theta_R, \theta_0', \theta_R'$ are real-valued. In the following we denote by ${\mathbb{C}}^{n \times n}$, $n\in{\mathbb{N}}$, the set of $n \times n$ matrices with complex-valued entries, and by $I_n$ the identity matrix in ${\mathbb{C}}^n$. Let $A= \big[A_{j,k}\big]_{1\leq j,k\leq 2}\in {\mathbb{C}}^{4 \times 4}$, with $A_{j,k}\in {\mathbb{C}}^{2 \times 2}$, $1\leq j,k\leq 2$, and $L\in {\mathbb{C}}^{2 \times 2}$, chosen such that $\ker (A_{1,1} +A_{1,2} L)=\{0\}$; that is, $(A_{1,1}+A_{1,2} L)$ is invertible in ${\mathbb{C}}^2$. Define for such $A$ (cf., e.g., \cite{KS74}), \begin{equation} M_A(L)=(A_{2,1}+A_{2,2} L)(A_{1,1}+A_{1,2} L)^{-1}, \label{4.1} \end{equation} and observe that \begin{align} & M_{I_{4}} (L) = L, \\ & M_{AB}(L) = M_A(M_B(L)), \label{4.2} \\ & M_{A^{-1}} (M_A(L)) = L = M_A(M_{A^{-1}} (L)), \quad A \, \text{ invertible,} \label{4.2a} \\ & M_A(L)=M_{AB^{-1}}(M_B(L)), \label{4.2b} \end{align} whenever the right-hand sides (and hence the left-hand sides) in \eqref{4.2}--\eqref{4.2b} exist. \begin{theorem} \label{t4.1} Assume that $\theta_0, \theta_R, \theta_0^{\prime},\theta_R^{\prime}, \delta_0,\delta_R,\delta_0^\prime,\delta_R^\prime\in S_{2 \pi}$, $\delta_0^\prime-\delta_0\ne 0 \, \text{\rm mod} (\pi)$, $\delta_R^\prime-\delta_R \ne 0 \, \text{\rm mod} (\pi)$, and that $z\in{\mathbb{C}}\backslash\big(\sigma(H_{\theta_0,\theta_R})\cup\sigma(H_{\delta_0,\delta_R})\big)$. Then, with $\ensuremath{S_{\te_0,\te_R}}$ defined as in \eqref{2.14}, \begin{align} \begin{split} \ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}}(z) &= \big(S_{\delta_0^\prime-\delta_0,\delta_R^\prime-\delta_R}\big)^{-1} \big[S_{\delta_0^\prime-\theta_0^\prime,\delta_R^\prime-\theta_R^\prime}+ S_{\theta_0^\prime-\delta_0,\theta_R^\prime-\delta_R}\ensuremath{\La_{\de_0,\de_R}^{\de_0^{\prime},\de_R^{\prime}}}(z) \big] \label{4.3} \\ & \quad \times \big[S_{\delta_0^\prime-\theta_0,\delta_R^\prime-\theta_R}+ S_{\theta_0-\delta_0,\theta_R-\delta_R}\ensuremath{\La_{\de_0,\de_R}^{\de_0^{\prime},\de_R^{\prime}}}(z) \big]^{-1}S_{\delta_0^\prime-\delta_0,\delta_R^\prime-\delta_R}. \end{split} \end{align} \end{theorem} \begin{proof} Assume $\theta_0, \theta_R, \theta_0^{\prime},\theta_R^{\prime}\in S_{2 \pi}$, $z\in{\mathbb{C}}\backslash\sigma(\ensuremath{H_{\te_0,\te_R}})$, and let $\psi_j(z,\cdot)$, $j=1,2$, denote a basis for the solutions of \eqref{2.3}. Then, with $\ensuremath{C_{\te_0,\te_R}}$ and $\ensuremath{S_{\te_0,\te_R}}$ as defined in \eqref{2.14}, equation \eqref{2.13} yields \begin{align}\label{4.4} &\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z) \bigg(\ensuremath{C_{\te_0,\te_R}}\begin{bmatrix} \psi_1(z,0)&\psi_2(z,0)\\ \psi_1(z,R)&\psi_2(z,R)\end{bmatrix}+ \ensuremath{S_{\te_0,\te_R}}\begin{bmatrix} \psi'_1(z,0)&\psi'_2(z,0)\\ -\psi'_1(z,R)&-\psi'_2(z,R)\end{bmatrix}\bigg)\notag \\ & \quad =\bigg(\ensuremath{C_{\te_0^{\prime},\te_R^{\prime}}}\begin{bmatrix} \psi_1(z,0)&\psi_2(z,0)\\ \psi_1(z,R)&\psi_2(z,R)\end{bmatrix}+ \ensuremath{S_{\te_0^{\prime},\te_R^{\prime}}}\begin{bmatrix} \psi'_1(z,0)&\psi'_2(z,0)\\ -\psi'_1(z,R)&-\psi'_2(z,R)\end{bmatrix}\bigg). \end{align} From Remark~\ref{r2.6}, recall, for $z\in{\mathbb{C}}\backslash\sigma(H_{0,0})$, that $\Lambda_{D,N}(z)=\ensuremath{\La_{0,0}^{\frac{\pi}{2},\frac{\pi}{2}}}(z)=\Lambda_{0,0}(z)$ and note that \eqref{2.32} then yields \begin{equation}\label{4.5} \Lambda_{0,0}(z) = \begin{bmatrix} \psi'_1(z,0)&\psi'_2(z,0)\\ -\psi'_1(z,R)&-\psi'_2(z,R) \end{bmatrix} \begin{bmatrix} \psi_1(z,0)&\psi_2(z,0)\\ \psi_1(z,R)&\psi_2(z,R)\end{bmatrix}^{-1}. \end{equation} Then, with $\ensuremath{C_{\te_0,\te_R}}$ defined as in \eqref{2.14}, and with $z\in{\mathbb{C}}\big\backslash\big(\sigma(\ensuremath{H_{\te_0,\te_R}})\cup \sigma(H_{0,0})\big)$, \eqref{4.4} can be written as \begin{equation} \label{4.6} \ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}}(z) =\big[\ensuremath{C_{\te_0^{\prime},\te_R^{\prime}}}+\ensuremath{S_{\te_0^{\prime},\te_R^{\prime}}} \Lambda_{0,0}(z)\big] \big[\ensuremath{C_{\te_0,\te_R}}+\ensuremath{S_{\te_0,\te_R}} \Lambda_{0,0}(z)\big]^{-1}. \end{equation} Next, assume that $\theta_0, \theta_R, \theta_0^{\prime},\theta_R^{\prime}, \delta_0,\delta_R,\delta_0^\prime,\delta_R^\prime\in S_{2 \pi}$, and let $A, B\in {\mathbb{C}}^{4 \times 4}$ be defined by \begin{equation}\label{4.7} A= \begin{bmatrix} \ensuremath{C_{\te_0,\te_R}} & \ensuremath{S_{\te_0,\te_R}} \\ C_{\theta_0^\prime,\theta_R^\prime} & S_{\theta_0^\prime,\theta_R^\prime}\end{bmatrix},\quad B=\begin{bmatrix} C_{\delta_0,\delta_R} & S_{\delta_0,\delta_R}\\ C_{\delta_0^\prime,\delta_R^\prime} & S_{\delta_0^\prime,\delta_R^\prime}\end{bmatrix}. \end{equation} Then, by \eqref{4.1}, and \eqref{4.6}, \begin{equation}\label{4.8} \ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}}(z) = M_A(\Lambda_{0,0}(z)), \qquad \ensuremath{\La_{\de_0,\de_R}^{\de_0^{\prime},\de_R^{\prime}}}(z) = M_B(\Lambda_{0,0}(z)) \end{equation} for $z\in{\mathbb{C}}\big\backslash\big(\sigma(\ensuremath{H_{\te_0,\te_R}}) \cup\sigma(H_{\delta_0,\delta_R})\cup \sigma(H_{0,0})\big)$. If additionally, one assumes that $\delta_0^\prime-\delta_0\ne 0 \, \text{\rm mod} (\pi)$, $\delta_R^\prime-\delta_R \ne 0 \, \text{\rm mod} (\pi)$, then \begin{equation}\label{4.9} AB^{-1}= \begin{bmatrix} \big(S_{\delta_0^\prime-\delta_0,\delta_R^\prime-\delta_R}\big)^{-1}S_{\delta_0^\prime-\theta_0,\delta_R^\prime-\theta_R} & \big(S_{\delta_0^\prime-\delta_0,\delta_R^\prime-\delta_R}\big)^{-1}S_{\theta_0-\delta_0,\theta_R-\delta_R}\\ \big(S_{\delta_0^\prime-\delta_0,\delta_R^\prime-\delta_R}\big)^{-1}S_{\delta_0^\prime-\theta_0^\prime,\delta_R^\prime-\theta_R^\prime} & \big(S_{\delta_0^\prime-\delta_0,\delta_R^\prime-\delta_R}\big)^{-1}S_{\theta_0^\prime-\delta_0,\theta_R^\prime-\delta_R} \end{bmatrix}, \end{equation} and \eqref{4.3} then follows from \eqref{4.1}, \eqref{4.2}, \eqref{4.8} and \eqref{4.9}, given that \begin{equation}\label{4.10} \ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}}(z) = M_A(\Lambda_{0,0}(z))=M_{AB^{-1}}(M_B(\Lambda_{0,0}(z))=M_{AB^{-1}}(\ensuremath{\La_{\de_0,\de_R}^{\de_0^{\prime},\de_R^{\prime}}}(z)). \end{equation} By meromorphic continuation, \eqref{4.3} holds for $z\in{\mathbb{C}}\big\backslash\big(\sigma(\ensuremath{H_{\te_0,\te_R}})\cup\sigma(H_{\delta_0,\delta_R})\big)$. \end{proof} If $\ensuremath{H_{\te_0,\te_R}}$ and $H_{\delta_0,\delta_R}$ are self-adjoint, then \eqref{4.3} holds for $z\in{\mathbb{C}}\backslash{\mathbb{R}}$. \begin{remark} \label{r4.2} In the special case of \eqref{4.3} which relates two generalized Dirichlet-to-Neumann maps one obtains \begin{align}\label{4.11} \ensuremath{\La_{\te_0,\te_R}}(z)&=\big[-S_{\theta_0-\delta_0,\theta_R-\delta_R} + C_{\theta_0-\delta_0,\theta_R-\delta_R} \ensuremath{\La_{\de_0,\de_R}}(z)\big]\notag\\ & \quad \times\big[C_{\theta_0-\delta_0,\theta_R-\delta_R}+ S_{\theta_0-\delta_0,\theta_R-\delta_R} \ensuremath{\La_{\de_0,\de_R}}(z)\big]^{-1}, \\ & \hspace*{-.6cm} \theta_0, \theta_R, \delta_0,\delta_R\in S_{2 \pi}, \; z\in{\mathbb{C}}\backslash\big(\sigma(H_{\theta_0,\theta_R})\cup\sigma(H_{\delta_0,\delta_R})\big). \notag \end{align} \end{remark} The following reformulation of \eqref{4.3} (motivated by the form of \eqref{3.33aa} in Theorem \ref{t3.5}) will be crucial in the proof of Theorem \ref{t4.6}. \begin{corollary} \label{c4.3} Assume that $\theta_0, \theta_R, \theta_0^{\prime},\theta_R^{\prime}, \delta_0,\delta_R,\delta_0^\prime,\delta_R^\prime\in S_{2 \pi}$, $\theta_0^\prime-\theta_0\ne 0 \, \text{\rm mod} (\pi)$, $\theta_R^\prime-\theta_R \ne 0 \, \text{\rm mod} (\pi)$, $\delta_0^\prime-\delta_0\ne 0 \, \text{\rm mod} (\pi)$, $\delta_R^\prime-\delta_R \ne 0 \, \text{\rm mod} (\pi)$, and that $z\in{\mathbb{C}}\backslash\big(\sigma(H_{\theta_0,\theta_R})\cup\sigma(H_{\delta_0,\delta_R})\big)$. Then \begin{align} \label{4.11a} & \ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}}(z) S_{\theta_0'-\theta_0,\theta_R'-\theta_R} \notag \\ & \quad = \Big[S_{\delta_0^\prime-\theta_0^\prime,\delta_R^\prime-\theta_R^\prime} + \big(S_{\delta_0^\prime-\delta_0,\delta_R^\prime-\delta_R}\big)^{-1} S_{\theta_0^\prime-\delta_0,\theta_R^\prime-\delta_R}\ensuremath{\La_{\de_0,\de_R}^{\de_0^{\prime},\de_R^{\prime}}}(z) S_{\delta_0^\prime-\delta_0,\delta_R^\prime-\delta_R}\Big] \notag\\ & \qquad \, \times \Big[\big(S_{\theta_0'-\theta_0,\theta_R'-\theta_R}\big)^{-1} S_{\delta_0^\prime-\theta_0,\delta_R^\prime-\theta_R} + \big(S_{\theta_0'-\theta_0,\theta_R'-\theta_R}\big)^{-1} \big(S_{\delta_0^\prime-\delta_0,\delta_R^\prime-\delta_R}\big)^{-1} \notag \\ & \hspace*{4.5cm} \times S_{\theta_0-\delta_0,\theta_R-\delta_R}\ensuremath{\La_{\de_0,\de_R}^{\de_0^{\prime},\de_R^{\prime}}}(z) S_{\delta_0'-\delta_0,\delta_R'-\delta_R}\Big]^{-1}. \end{align} \end{corollary} \begin{proof} This follows from multiplying \eqref{4.3} by $S_{\theta_0'-\theta_0,\theta_R'-\theta_R}$ from the right, inserting the terms $S_{\delta_0^\prime-\delta_0,\delta_R^\prime-\delta_R} \big(S_{\delta_0^\prime-\delta_0,\delta_R^\prime-\delta_R}\big)^{-1}$ to the right of $\ensuremath{\La_{\de_0,\de_R}^{\de_0^{\prime},\de_R^{\prime}}}(z)$ twice, and then algebraically manipulate the various terms in the equation resulting from these insertions into \eqref{4.3} (repeatedly using $(ST)^{-1} = T^{-1}S^{-1}$, etc.). \end{proof} Since the self-adjoint case is the principal focus for the remainder of this section, we now recall some additional pertinent facts from \cite{KS74} (see also \cite[Sect.\ 6]{GT00}) on linear fractional transformations of matrices. Defining \begin{equation}\label{4.12} J_{4}=\begin{bmatrix} 0 & -I_2\\ I_2 & 0\end{bmatrix}, \end{equation} and \begin{equation}\label{4.13} {\mathcal A}_{4}=\{ A\in{\mathbb{C}}^{4 \times 4} \, | \, A^*J_{4}A=J_{4}\}, \end{equation} representing $A\in {\mathbb{C}}^{4 \times 4}$ by \begin{equation}\label{4.14} A= \begin{bmatrix} A_{1,1} & A_{1,2}\\ A_{2,1} & A_{2,2}\end{bmatrix}, \quad A_{p,q}\in {\mathbb{C}}^{2 \times 2},\quad 1\leq p,q\leq 2, \end{equation} the condition $A^*J_{4}A=J_{4}$ in \eqref{4.13} is equivalent to \begin{align} \begin{split} \label{4.15} & A^*_{1,1}A_{2,1}=A^*_{2,1}A_{1,1},\quad A^*_{2,2}A_{1,2}=A^*_{1,2}A_{2,2}, \\ & A^*_{2,2}A_{1,1}-A^*_{1,2}A_{2,1} = I_2 = A^*_{1,1}A_{2,2} - A^*_{2,1}A_{1,2}, \end{split} \end{align} or equivalently, to \begin{equation}\label{4.16} \begin{bmatrix}A^*_{2,2} & -A^*_{1,2}\\-A^*_{2,1}& A^*_{1,1}\end{bmatrix} \begin{bmatrix} A_{1,1} & A_{1,2}\\ A_{2,1} & A_{2,2}\end{bmatrix} = I_{4}. \end{equation} Since left inverses in ${\mathbb{C}}^{4 \times 4}$ are also right inverses, \eqref{4.16} implies \begin{equation}\label{4.17} \begin{bmatrix} A_{1,1} & A_{1,2}\\ A_{2,1} & A_{2,2}\end{bmatrix} \begin{bmatrix} A^*_{2,2} & -A^*_{1,2}\\ -A^*_{2,1} & A^*_{1,1}\end{bmatrix} = I_{4}, \end{equation} that is, \begin{align} \begin{split} \label{4.18} & A_{1,1}A^*_{1,2}=A_{1,2} A^*_{1,1},\quad A_{2,2}A^*_{2,1}=A_{2,1}A^*_{2,2}, \\ & A_{2,2}A^*_{1,1}-A_{2,1}A^*_{1,2}=I_{2} =A_{1,1}A^*_{2,2}-A_{1,2} A^*_{2,1}, \end{split} \end{align} or equivalently, \begin{equation}\label{4.19} AJ_{4}A^*=J_{4}. \end{equation} In particular, \begin{equation}\label{4.20} A\in {\mathcal A}_{4} \, \text{ if and only if } \, A^{-1}\in {\mathcal A}_{4}. \end{equation} At this point we turn to the particularly important special self-adjoint case where $V$ and $\theta_0, \theta_R, \theta_0', \theta_R'$ are real-valued. In this case we will now prove that $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}}(z) S_{\theta_0'-\theta_0,\theta_R'-\theta_R}$ is a $2 \times 2$ matrix-valued Herglotz function. But first we note the following result: \begin{lemma} \label{l4.4} Assume that $\theta_0, \theta_R, \theta_0^{\prime},\theta_R^{\prime}, \delta_0,\delta_R,\delta_0^\prime,\delta_R^\prime\in [0,2 \pi)$, $\theta_0^\prime-\theta_0\ne 0 \, \text{\rm mod} (\pi)$, $\theta_R^\prime-\theta_R \ne 0 \, \text{\rm mod} (\pi)$, $\delta_0^\prime-\delta_0\ne 0 \, \text{\rm mod} (\pi)$, $\delta_R^\prime-\delta_R \ne 0 \, \text{\rm mod} (\pi)$, and introduce in accordance with \eqref{4.11a}, \begin{align} \begin{split} \label{4.22} & A(\theta, \delta) = \big[ A(\theta, \delta)_{j,k}\big]_{1 \leq j,k \leq 2} \in {\mathbb{C}}^{4 \times 4}, \\ & A(\theta, \delta)_{1,1} = \big(S_{\theta_0'-\theta_0,\theta_R'-\theta_R}\big)^{-1} S_{\delta_0'-\theta_0,\delta_R'-\theta_R}, \\ & A(\theta, \delta)_{1,2} = \big(S_{\theta_0'-\theta_0,\theta_R'-\theta_R}\big)^{-1} \big(S_{\delta_0'-\delta_0,\delta_R'-\delta_R}\big)^{-1} S_{\theta_0-\delta_0,\theta_R-\delta_R}, \\ & A(\theta, \delta)_{2,1} = S_{\delta_0'-\theta_0',\delta_R'-\theta_R'}, \\ & A(\theta, \delta)_{2,2} = \big(S_{\delta_0'-\delta_0,\delta_R'-\delta_R}\big)^{-1} S_{\theta_0'-\delta_0,\theta_R'-\delta_R}. \end{split} \end{align} Then \begin{equation} A(\theta, \delta) \in {\mathcal A}_4. \label{4.23} \end{equation} \end{lemma} \begin{proof} Since according to \eqref{2.14} $S_{\alpha,\beta}$ are $2 \times 2$ diagonal matrices, an entirely straightforward (though, admittedly, rather tedious) explicit computation shows that the block matrix entries of $A_{\theta, \delta}$ in \eqref{4.22} satisfy the relations in \eqref{4.15}. \end{proof} We denote by ${\mathbb{C}}_+$ the open complex upper half-plane and abbreviate $\Im(L)=(L - L^*)/(2i)$ for $L \in{\mathbb{C}}^{n\times n}$, $n\in{\mathbb{N}}$. In addition, $d \|\Sigma\|_{{\mathbb{C}}^2}$ will denote the total variation of the $2\times 2$ matrix-valued measure $d \Sigma$ below in \eqref{5.2}. We recall that $M(\cdot)$ is called an $n \times n$ matrix-valued Herglotz function if it is analytic on ${\mathbb{C}}_+$ and $\Im(M(z)) \geq 0$ for all $z\in{\mathbb{C}}_+$. In this context we also recall the following result: \begin{lemma} \label{l4.5} Assume that $A=\big[A_{j,k}\big]_{1\leq j,k\leq 2}\in {\mathcal A}_4$ and $L \in {\mathbb{C}}^{2 \times 2}$. Then \begin{equation} \Im(L) > 0 \, \text{ implies } \, \ker (A_{1,1} + A_{1,2} L)=\{0\} \label{4.24} \end{equation} and $M_A(L)=(A_{2,1}+A_{2,2} L)(A_{1,1}+A_{1,2} L)^{-1}$ $($defined according to \eqref{4.1}$)$ satisfies \begin{equation} \Im(M_A) = \big((A_{2,1} + A_{2,2} L)^{-1}\big)^* \Im(L) \label{4.25} (A_{2,1} + A_{2,2} L)^{-1} > 0. \end{equation} In particular, if $M(\cdot)$ is a $2 \times 2$ matrix-valued Herglotz function satisfying \begin{equation} \Im(M(z)) > 0, \quad z\in{\mathbb{C}}_+, \label{4.26} \end{equation} then $M_A(\cdot)$ $($in obvious notation defined according to \eqref{4.1} with $L$ replaced by $M(\cdot)$$)$ is a a $2 \times 2$ matrix-valued Herglotz function satisfying \begin{equation} \Im(M_A(z)) > 0, \quad z\in{\mathbb{C}}_+. \label{4.27} \end{equation} \end{lemma} \begin{proof} This is the special finite-dimensional case of \cite[Theorem\ 6.4]{GT00}. \end{proof} Now we are in position to prove the fundamental Herglotz property of the matrix $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (\cdot) S_{\theta_0'-\theta_0,\theta_R'-\theta_R}$ in the case where $\ensuremath{H_{\te_0,\te_R}}$ is self-adjoint. \begin{theorem} \label{t4.6} Let $\theta_0, \theta_R, \theta_0', \theta_R' \in [0,2 \pi)$, $\theta_0^\prime-\theta_0\ne 0 \, \text{\rm mod} (\pi)$, $\theta_R^\prime-\theta_R \ne 0 \, \text{\rm mod} (\pi)$, $z\in{\mathbb{C}}\backslash\sigma(\ensuremath{H_{\te_0,\te_R}})$, and $\ensuremath{H_{\te_0,\te_R}}$ be defined as in \eqref{2.2a}. In addition, suppose that $V$ is real-valued $($and hence $\ensuremath{H_{\te_0,\te_R}}$ is self-adjoint\,$)$. Then $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (\cdot) S_{\theta_0'-\theta_0,\theta_R'-\theta_R}$ is a $2\times 2$ matrix-valued Herglotz function admitting the representation \begin{align} & \ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z) S_{\theta_0'-\theta_0,\theta_R'-\theta_R}=\Xi_{\theta_0,\theta_R}^{\theta_0', \theta_R'} + \ \int_{{\mathbb{R}}} d\Sigma_{\theta_0,\theta_R}^{\theta_0', \theta_R'} (\lambda)\bigg(\frac{1}{\lambda -z}-\frac{\lambda} {1+\lambda^2}\bigg), \label{5.1} \\ & \hspace*{7.75cm} z\in{\mathbb{C}}\backslash\sigma(\ensuremath{H_{\te_0,\te_R}}), \notag \\ & \Xi_{\theta_0,\theta_R}^{\theta_0', \theta_R'} = \Big(\Xi_{\theta_0,\theta_R}^{\theta_0', \theta_R'}\Big)^* \in{\mathbb{C}}^{2\times 2}, \quad \int_{{\mathbb{R}}} \frac{d\big\|\Sigma_{\theta_0,\theta_R}^{\theta_0', \theta_R'} (\lambda)\big\|_{{\mathbb{C}}^2}}{1+\lambda^2} <\infty, \label{5.2} \end{align} where \begin{align} \begin{split} \Sigma_{\theta_0,\theta_R}^{\theta_0', \theta_R'} ((\lambda_1,\lambda_2]) = \frac{1}{\pi}\lim_{\delta\downarrow 0}\lim_{\varepsilon\downarrow 0}\int^{\lambda_2+\delta}_{\lambda_1+\delta}d\lambda \, \Im\Big(\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (\lambda +i\varepsilon)\Big),& \\ \lambda_1, \lambda_2 \in{\mathbb{R}}, \; \lambda_1<\lambda_2.& \label{5.3} \end{split} \end{align} In addition, \begin{equation} \Im\Big(\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z) S_{\theta_0'-\theta_0,\theta_R'-\theta_R}\Big) > 0, \quad z\in{\mathbb{C}}_+. \label{5.3a} \end{equation} \end{theorem} \begin{proof} Without loss of generality, we take $z\in{\mathbb{C}}_+$. An analytic continuation with respect to $z$ then extends the result \eqref{5.1} to $z\in{\mathbb{C}}\backslash\sigma(\ensuremath{H_{\te_0,\te_R}})$. We will first prove the Herglotz property for $\Lambda_{\pi/2,\pi/2}$ and then use a special case of the linear fractional transformation \eqref{4.11a} to conclude that $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}}(\cdot) S_{\theta_0'-\theta_0,\theta_R'-\theta_R}$ is a $2 \times 2$ matrix-valued Herglotz function for any $\theta_0, \theta_R, \theta_0', \theta_R' \in [0,2 \pi)$. Our point of departure is formula \eqref{3.33a} in the special case $\theta_0 = \theta_R = \pi/2$, that is, \begin{equation} \Lambda_{\pi/2,\pi/2} (z) = \gamma_{0,0} \big[\gamma_{0,0} (H_{\pi/2,\pi/2}- \overline{z})^{-1}]^*, \quad z \in {\mathbb{C}}_+, \label{5.4} \end{equation} noticing that \begin{equation} \widehat \gamma_{\pi/2,\pi/2} = \gamma_{\pi,\pi} = - \gamma_{0,0} \end{equation} by \eqref{2.2A}. First, we slightly change the definition of $\gamma_{0,0}$ by introducing \begin{equation} \label{5.5} \widetilde \gamma_{0,0} \colon \begin{cases} H^1((0,R)) \rightarrow {\mathbb{C}}^2, \\ u \mapsto \begin{bmatrix} u(0) \\ u(R) \end{bmatrix}, \end{cases} \quad \widetilde \gamma_{0,0} \in {\mathcal B}\big(H^1((0,R)), {\mathbb{C}}^2\big), \end{equation} instead. One notes that $\widetilde \gamma_{0,0}$ is well-defined in the following sense: Any $u \in H^1((0,R))$ has a representative in its equivalence class of Lebesgue measurable and square integrable elements, again denoted by $u$ for simplicity, that is absolutely continuous on $[0,R]$. In fact, by a standard Sobolev embedding result, one has $H^1(0,R))\hookrightarrow C^{1/2}((0,R)) = C^{1/2}([0,R])$. In particular, the limits $\lim_{x\downarrow 0} u(x) = u(0)$ and $\lim_{x\uparrow R} u(x) = u(R)$ are well-defined for this representative $u$. Next, using \begin{equation} H^1((0,R)) \hookrightarrow L^2((0,R); dx)) \hookrightarrow H^1((0,R))^*, \label{5.6} \end{equation} one infers that (cf.\ also Remark \ref{r3.7}) \begin{equation} \label{5.7} (\widetilde \gamma_{0,0})^* \colon \begin{cases} {\mathbb{C}}^2 \rightarrow H^1((0,R))^*, \\ \begin{bmatrix} v_1 \\ v_2 \end{bmatrix} \mapsto v_1 \delta_0 + v_2 \delta_R, \end{cases} \quad (\widetilde \gamma_{0,0})^* \in {\mathcal B}\big({\mathbb{C}}^2, H^1((0,R))^*\big), \end{equation} where (cf.\ also \cite[Example\ 2]{KS95}) \begin{align} & \delta_0, \delta_R \in H^1((0,R))^*, \notag \\ & \delta_0 (u) = {}_{H^1((0,R))}\big\langle \overline u, \delta_0 \big\rangle_{H^1((0,R))^*} = u(0), \label{5.8} \\ & \delta_R (u) = {}_{H^1((0,R))}\big\langle \overline u, \delta_R \big\rangle_{H^1((0,R))^*} = u(R), \quad u \in H^1((0,R)), \notag \end{align} with ${}_{H^1((0,R))}\langle \, \cdot \, , \cdot \, \rangle_{H^1((0,R))^*}$ denoting the duality pairing between $H^1((0,R))$ and $H^1((0,R))^*$ (linear in the second argument and antilinear in the first, see, e.g., \cite[Sect.\ 2]{GM09a} for more details). Indeed, \eqref{5.8} follows from \begin{align} & u(0) v_1 + u(R) v_2 = (\widetilde \gamma_{0,0} \overline u,v)_{{\mathbb{C}}^2} = {}_{H^1((0,R))}\big\langle \overline u, (\widetilde \gamma_{0,0})^* v \big\rangle_{H^1((0,R))^*}, \\ & \hspace*{4.8cm} v = \begin{bmatrix} v_1 \\ v_2 \end{bmatrix} \in {\mathbb{C}}^2, \quad u \in H^1((0,R)). \label{5.9} \end{align} Next, still following the material discussed in \cite[Sect.\ 2]{GM09a}, we extend the operator $H_{\pi/2,\pi/2}$ in $L^2((0,R); dx)$, \begin{equation} (H_{\pi/2,\pi/2} -z I ) \colon \text{\rm{dom}}\big(H_{\pi/2,\pi/2}\big) \to L^2((0,R); dx), \quad z \in z\in{\mathbb{C}}_+, \label{5.10} \end{equation} where $\text{\rm{dom}}\big(H_{\pi/2,\pi/2}\big) \hookrightarrow L^2((0,R); dx)$, to its extension $\widetilde {H_{\pi/2,\pi/2}}$, which maps $H^1((0,R))$ boundedly into $H^1((0,R))^*$, \begin{equation} \widetilde {H_{\pi/2,\pi/2}} \in {\mathcal B}\big(H^1((0,R)), H^1((0,R))^*\big), \label{5.11} \end{equation} such that (with $\widetilde I: H^1((0,R)) \hookrightarrow H^1((0,R))^*$ the continuous embedding operator) \begin{align} & \big(\widetilde {H_{\pi/2,\pi/2}}+\widetilde I\,\big)\in{\mathcal B}\big(H^1((0,R)), H^1((0,R))^*\big) \\ & \, \text{and }\big(\widetilde {H_{\pi/2,\pi/2}} + \widetilde I\,\big) \colon H^1((0,R)) \to H^1((0,R))^* \, \text{ is unitary.} \label{B.37} \end{align} In addition (cf.\ \eqref{2.2f}), \begin{align} & (H_{\pi/2,\pi/2}+I)^{1/2}\in{\mathcal B}\big(H^1((0,R)), L^2((0,R);dx)\big) \\ & \, \text{and } (H_{\pi/2,\pi/2} + I)^{1/2} \colon H^1((0,R)) \to L^2((0,R);dx) \, \text{ is unitary} \label{B.40} \end{align} (cf.\ \cite[Sect.\ 2]{GM09a}). Moreover, $\widetilde {H_{\pi/2,\pi/2}}$ is self-adjoint, \begin{equation} \Big(\widetilde {H_{\pi/2,\pi/2}}\Big)^* = \widetilde {H_{\pi/2,\pi/2}}, \label{5.12} \end{equation} in the sense that \begin{align} \begin{split} {}_{H^1((0,R))}\big\langle w_1, \widetilde {H_{\pi/2,\pi/2}} w_2 \big\rangle_{H^1((0,R))^*} = \overline{{}_{H^1((0,R))}\big\langle w_2, \widetilde {H_{\pi/2,\pi/2}} w_1 \big\rangle_{H^1((0,R))^*}}&, \\ \quad w_1, w_2 \in H^1((0,R))& \label{5.12a} \end{split} \end{align} (again we refer to \cite[Sect.\ 2]{GM09a} for more details). In addition, \begin{equation} \Big(\widetilde {H_{\pi/2,\pi/2}} -z \widetilde I \Big)^{-1} \colon H^1((0,R))^* \to H^1((0,R)), \quad z \in{\mathbb{C}}_+, \label{5.13} \end{equation} and hence, \begin{align} \begin{split} \Big(\Big(\widetilde {H_{\pi/2,\pi/2}} -z \widetilde I \Big)^{-1} w\Big)(x) = {}_{H^1((0,R))}\big\langle \overline{G_{\pi/2,\pi/2} (z,x,\cdot)}, w \big\rangle_{H^1((0,R))^*},& \\ w \in H^1((0,R))^*, \; z \in{\mathbb{C}}_+,& \label{5.14} \end{split} \end{align} using the fact that \begin{equation} G_{\pi/2,\pi/2} (z,x,\cdot) \in H^1((0,R)), \quad x \in{\mathbb{R}}, \; z\in{\mathbb{C}}\backslash\sigma(H_{\pi/2,\pi/2}). \end{equation} By \eqref{2.5a} and \eqref{2.5b}, the Wronskian $W$ is of the form \begin{align} \begin{split} W(z) & = W(u_{+,\pi/2}(z,\cdot), u_{-,\pi/2}(z,\cdot)) \\ & = - u'_{+,\pi/2}(z,0) u_{-,\pi/2}(z,0) = u_{+,\pi/2}(z,R) u'_{-,\pi/2}(z,R). \label{5.15} \end{split} \end{align} Using \eqref{3.32}, one computes \begin{align} & \Big(\Big(\widetilde {H_{\pi/2,\pi/2}} -z \widetilde I \Big)^{-1} (\widetilde \gamma_{0,0})^* v \Big)(x) \notag \\ & \quad = \frac{1}{W(z)} {}_{H^1((0,R))}\big\langle \overline{G_{\pi/2,\pi/2} (z,x,\cdot)}, [v_1 \delta_0 + v_2 \delta_R] \big\rangle_{H^1((0,R))^*} \notag \\ & \quad = \frac{1}{W(z)} \big( u_{-,\pi/2}(z,x) u_{+,\pi/2} (z,R) v_2 + u_{+,\pi/2}(z,x) u_{-,\pi/2} (z,0) v_1 \big), \label{5.16} \\ & \hspace*{7.6cm} v = [v_1 \;\, v_2]^\top \in {\mathbb{C}}^2, \notag \end{align} and hence \begin{align} & \widetilde \gamma_{0,0} \Big(\widetilde {H_{\pi/2,\pi/2}} - z \widetilde I \Big)^{-1} (\widetilde \gamma_{0,0})^* v \notag \\ & \quad = \frac{1}{W(z)}\begin{bmatrix} u_{-,\pi/2}(z,0) u_{+,\pi/2}(z,R) v_2 + u_{-,\pi/2}(z,0) v_1 \\[1mm] u_{+,\pi/2}(z,R) v_2 + u_{+,\pi/2}(z,R) u_{-,\pi/2}(z,0) v_1 \end{bmatrix} \notag \\[1mm] & \quad = \frac{1}{W(z)} \begin{bmatrix} u_{-,\pi/2}(z,0) & u_{-,\pi/2}(z,0) u_{+,\pi/2}(z,R) \\[1mm] u_{-,\pi/2}(z,0) u_{+,\pi/2}(z,R) & u_{+,\pi/2}(z,R) \end{bmatrix} \begin{bmatrix} v_1 \\[1mm] v_2 \end{bmatrix} \notag \\[1mm] & \quad = \Lambda_{\pi/2,\pi/2} (z) v, \quad v = [v_1 \;\, v_2]^\top \in {\mathbb{C}}^2, \label{5.17} \end{align} inserting \eqref{5.15} for $W(\cdot)$. Consequently, one has \begin{equation} \Lambda_{\pi/2,\pi/2} (z) = \widetilde \gamma_{0,0} \Big(\widetilde {H_{\pi/2,\pi/2}} - z \widetilde I \Big)^{-1} (\widetilde \gamma_{0,0})^*, \quad z\in{\mathbb{C}}_+, \label{5.18} \end{equation} and also \begin{equation} \Lambda_{\pi/2,\pi/2} (z)^* = \Lambda_{\pi/2,\pi/2} (\overline z), \quad z\in{\mathbb{C}}_+. \label{5.19} \end{equation} In particular, \begin{align} & \big(v,\Im \big(\Lambda_{\pi/2,\pi/2} (z)\big) v\big)_{{\mathbb{C}}^2} = \frac{1}{2 i} \big(v, \big(\Lambda_{\pi/2,\pi/2} (z) - \Lambda_{\pi/2,\pi/2} (z)^*\big) v\big)_{{\mathbb{C}}^2} \notag \\ & \quad = \frac{1}{2 i} \Big(v, \widetilde \gamma_{0,0} \Big(\Big(\widetilde {H_{\pi/2,\pi/2}} - z \widetilde I \Big)^{-1} - \Big(\widetilde {H_{\pi/2,\pi/2}} -\overline{z} \widetilde I \Big)^{-1}\Big) (\widetilde \gamma_{0,0})^* v\Big)_{{\mathbb{C}}^2} \notag \\ & \quad = \Im(z) \Big(v, \widetilde \gamma_{0,0} \Big[\Big(\widetilde {H_{\pi/2,\pi/2}} - z \widetilde I \Big)^{-1} \widetilde I \Big(\widetilde {H_{\pi/2,\pi/2}} -\overline{z} \widetilde I \Big)^{-1}\Big] (\widetilde \gamma_{0,0})^* v\Big)_{{\mathbb{C}}^2} \notag \\ & \quad = \Im(z) \Big(v, \Big[\widetilde \gamma_{0,0} \Big(\widetilde {H_{\pi/2,\pi/2}} - z \widetilde I \Big)^{-1}\Big] \widetilde I \Big[\widetilde \gamma_{0,0} \Big(\widetilde {H_{\pi/2,\pi/2}} - z \widetilde I \Big)^{-1}\Big]^* v\Big)_{{\mathbb{C}}^2} \notag \\ & \quad = \Im(z)\, {}_{H^1((0,R))}\Big\langle \Big[\widetilde \gamma_{0,0} \Big(\widetilde {H_{\pi/2,\pi/2}} - z \widetilde I \Big)^{-1}\Big]^* v,\widetilde I \notag \\ & \hspace*{3.2cm} \times \Big[\widetilde \gamma_{0,0} \Big(\widetilde {H_{\pi/2,\pi/2}} - z \widetilde I \Big)^{-1}\Big]^* v\Big\rangle_{H^1((0,R))^*} \notag \\ & \quad = \Im(z) \big(\big[\gamma_{0,0} ({H_{\pi/2,\pi/2}} - z I)^{-1}\big]^* v, \big[\gamma_{0,0} ({H_{\pi/2,\pi/2}} - z I)^{-1}\big]^*v\big)_{L^2((0,R); dx)} \notag \\ & \quad = \Im(z) \big\|\big[\gamma_{0,0} ({H_{\pi/2,\pi/2}} - z I)^{-1}\big]^*v\big\|_{L^2((0,R); dx)}^2 \geq 0, \quad v \in {\mathbb{C}}^2, \; z\in{\mathbb{C}}_+, \label{5.20} \end{align} since the duality pairing ${}_{H^1((0,R))}\langle \, \cdot \, , \cdot \, \rangle_{H^1((0,R))^*}$ between the spaces $H^1((0,R))$ and $H^1((0,R))^*$ is compatible with the scalar product in $L^2((0,R); dx)$, that is, \begin{equation} {}_{H^1((0,R))}\langle w_1, \widetilde I w_2 \rangle_{H^1((0,R))^*} = (w_1, w_2)_{L^2((0,R); dx)}, \quad w_1, w_2 \in H^1((0,R)). \end{equation} In particular, \begin{align} \begin{split} \Im \big(\Lambda_{\pi/2,\pi/2} (z)\big) = \big[\gamma_{0,0} ({H_{\pi/2,\pi/2}} - z I)^{-1}\big] \big[\gamma_{0,0} ({H_{\pi/2,\pi/2}} - z I)^{-1}\big]^* \geq 0,& \label{5.20a} \\ z \in {\mathbb{C}}_+,& \end{split} \end{align} and thus, $\Lambda_{\pi/2,\pi/2} (\cdot)$ is a $2 \times 2$ matrix-valued Herglotz function. In fact, one can improve on \eqref{5.20a} to obtain \begin{equation} \Im \big(\Lambda_{\pi/2,\pi/2} (z)\big) > 0, \quad z\in{\mathbb{C}}_+, \label{5.21} \end{equation} since \begin{equation} \ker \big(\big[\gamma_{0,0} (H_{\pi/2,\pi/2} - z I)^{-1}\big]^*\big) = \{0\}, \quad z\in{\mathbb{C}}\backslash\sigma(H_{\pi/2,\pi/2}). \label{5.22} \end{equation} To prove \eqref{5.22}, one can argue as follows: Suppose that \begin{equation} [a_0 \;\, a_R]^\top \in \ker \big(\big[\gamma_{0,0} (H_{\pi/2,\pi/2} - z I)^{-1}\big]^*\big), \end{equation} then by \eqref{3.43}, \begin{align} & \big(\big[\gamma_{0,0} (H_{\pi/2,\pi/2} - z I)^{-1}\big]^* [a_0 \;\, a_R]^\top\big)(x) = \frac{1}{W(u_{+,\pi/2}(\overline z,\cdot), u_{-,\pi/2}(\overline z,\cdot))} \notag \\ & \qquad \times \big[a_0 u_{-,\pi/2}(\overline z,0) u_{+,\pi/2}(\overline z,x) + a_R u_{+,\pi/2}(\overline z,R) u_{-,\pi/2}(\overline z,x) \big] =0, \\ & \hspace*{7.65cm} z \in{\mathbb{C}}\backslash\sigma(H_{\pi/2,\pi/2}). \notag \end{align} Since by definition (cf.\ \eqref{2.5a}, \eqref{2.5b}), $u_{-,\pi/2}' (\overline z,0) = u_{+,\pi/2}' (\overline z,R) = 0$, one concludes that \begin{equation} u_{-,\pi/2} (\overline z,0) \neq 0, \quad u_{+,\pi/2} (\overline z,R) \neq 0. \end{equation} Moreover, since $W(u_{+,\pi/2}(\overline z,\cdot), u_{-,\pi/2}(\overline z,\cdot)) \neq 0$ for all $z \in{\mathbb{C}}\backslash\sigma(H_{\pi/2,\pi/2})$ (otherwise, $\overline z$ would be an eigenvalue of $H_{\pi/2,\pi/2}$), $u_{+,\pi/2}(\overline z,\cdot)$ and $u_{-,\pi/2}(\overline z,\cdot))$ are linearly independent, implying \begin{equation} a_0 = a_R = 0, \end{equation} and hence \eqref{5.22}. Next, using the notation introduced in \eqref{3.0}, and applying the linear fractional transformation \eqref{4.3} one can show that (with $z\in{\mathbb{C}}_+$) \begin{equation} \Lambda_{\pi/2,\pi/2} (z) \big[C_{\pi/2,\pi/2} + S_{\pi/2,\pi/2} \Lambda_{0,0} (z)\big] = \big[C_{\pi,\pi} + S_{\pi,\pi} \Lambda_{0,0} (z)\big], \end{equation} or equivalently, that \begin{equation} \Lambda_{\pi/2,\pi/2} (z) \Lambda_{0,0} (z) = - I_2, \end{equation} and hence, \begin{equation} \Lambda_{0,0}^{\pi/2,\pi/2} (z) = \Lambda_{0,0} (z) = -\big [\Lambda_{\pi/2,\pi/2} (z)\big]^{-1} = - \big[\Lambda_{\pi/2,\pi/2}^{\pi,\pi} (z)\big]^{-1}, \label{4.60} \end{equation} is a $2 \times 2$ matrix-valued Herglotz function too, satisfying \begin{equation} \Im \big(\Lambda_{0,0}^{\pi/2,\pi/2} (z)\big) > 0, \quad z\in{\mathbb{C}}_+. \label{4.61} \end{equation} The Herglotz property of $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (\cdot) S_{\theta_0'-\theta_0,\theta_R'-\theta_R}$, $\theta_0,\theta_R, \theta_0', \theta_R' \in [0,2 \pi)$ then follows again from the linear fractional transformation \eqref{4.11a} and from \eqref{4.27} in Lemma \ref{l4.5} upon identifying the $2\times 2$ block matrix $A = \big[A_{j,k}\big]_{1\leq j, k \leq 2}$ in Lemma \ref{l4.5} with $A(\theta,\delta) \in {\mathcal A}_4$ in the special case where \begin{align} \begin{split} & \theta_0, \theta_R, \theta_0', \theta_R' \in [0,2 \pi), \quad \theta_0^\prime-\theta_0\ne 0 \, \text{\rm mod} (\pi), \quad \theta_R^\prime-\theta_R \ne 0 \, \text{\rm mod} (\pi), \\ & \delta_0 = \delta_R =0, \quad \delta_0' = \delta_R' = \pi/2. \label{4.62} \end{split} \end{align} Given the Herglotz property of $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (\cdot) S_{\theta_0'-\theta_0,\theta_R'-\theta_R}$, one obtains as in \cite[Theorem\ 5.4]{GT00} the representation \begin{align} & \ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z) S_{\theta_0'-\theta_0,\theta_R'-\theta_R} =\Xi_{\theta_0,\theta_R}^{\theta_0', \theta_R'} + \Upsilon_{\theta_0,\theta_R}^{\theta_0', \theta_R'} z +\ \int_{{\mathbb{R}}} d\Sigma_{\theta_0,\theta_R}^{\theta_0', \theta_R'} (\lambda)\bigg(\frac{1}{\lambda -z}-\frac{\lambda} {1+\lambda^2}\bigg), \notag \\ & \hspace*{8.4cm} z\in{\mathbb{C}}\backslash\sigma(\ensuremath{H_{\te_0,\te_R}}), \label{5.26} \end{align} with $\Xi_{\theta_0,\theta_R}^{\theta_0', \theta_R'}$ and $\Sigma_{\theta_0,\theta_R}^{\theta_0', \theta_R'} (\cdot)$ as in \eqref{5.2} and \eqref{5.3}, and with $\Upsilon_{\theta_0,\theta_R}^{\theta_0', \theta_R'}$ satisfying \begin{equation} 0 \leq \Upsilon_{\theta_0,\theta_R}^{\theta_0', \theta_R'}\in{\mathbb{C}}^{2\times 2}. \label{5.27} \end{equation} Thus, to conclude the proof of \eqref{5.1}, it remains to prove that actually, $\Upsilon_{\theta_0,\theta_R}^{\theta_0', \theta_R'} = 0$. The latter fact is clear since \begin{equation} \ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z) \underset{\substack{|z|\to\infty\\ \Im(z^{1/2})>0}}{=} O(|z|^{1/2}), \end{equation} using the fact that by \eqref{4.3}, \begin{equation} \ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z) = C_{\theta_0'-\theta_0,\theta_R'-\theta_R} + S_{\theta_0'-\theta_0,\theta_R'-\theta_R} \ensuremath{\La_{\te_0,\te_R}} (z), \end{equation} and applying Lemma \ref{l3.4}. Finally, \eqref{5.3a} is a consequence of \eqref{4.61} and Lemma \ref{l4.5} with $A = A(\theta,\delta)$ chosen as in \eqref{4.62}. \end{proof} As a particular case of Theorem \ref{t4.6} where $\theta_0' = (\theta_0 + (\pi/2)) \, \text{\rm mod} (2\pi)$, $\theta_R' = (\theta_R + (\pi/2)) \, \text{\rm mod} (2\pi)$, one concludes that \begin{equation} \ensuremath{\La_{\te_0,\te_R}}(z) = \Lateq (z), \quad \theta_0,\theta_R\in [0,2 \pi), \; z\in{\mathbb{C}}\backslash \sigma(\ensuremath{H_{\te_0,\te_R}}), \label{4.76} \end{equation} is a $2 \times 2$ matrix-valued Herglotz function. \section{Krein-Type Resolvent Formulas} \label{s6} Krein-type resolvent formulas have been studied in a great variety of contexts, far too numerous to account for all in this paper. For instance, they are of fundamental importance in connection with the spectral and inverse spectral theory of ordinary and partial differential operators. Abstract versions of Krein-type resolvent formulas (see also the brief discussion at the end of our introduction), connected to boundary value spaces (boundary triples) and self-adjoint extensions of closed symmetric operators with equal (possibly infinite) deficiency spaces, have received enormous attention in the literature. In particular, we note that Robin-to-Dirichlet maps in the context of ordinary differential operators reduce to the celebrated (possibly, matrix-valued) Weyl--Titchmarsh function, the basic object of spectral analysis in this context. Since it is impossible to cover the literature in this paper, we refer the reader to the rather extensive recent bibliography in \cite{GM08}, \cite{GM09}, and \cite{GM10}. Here we just mention, for instance, \cite[Sect.\ 84]{AG81}, \cite{ABMN05}, \cite{AB09}, \cite{AP04}, \cite{AT03}, \cite{AT05}, \cite{BL07}, \cite{BMN08}, \cite{BMN02}, \cite{BGW09}, \cite{BHMNW09}, \cite{BM04}, \cite{BMNW08}, \cite{BGP08}, \cite{DHMS06}, \cite{DM91}, \cite{DM95}, \cite{GKMT01}, \cite{GMT98}, \cite{GMZ07}, \cite{GT00}, \cite[Ch.\ 3]{GG91}, \cite{Gr08a}, \cite[Ch.\ 13]{Gr09},\cite{KO77}, \cite{KO78}, \cite{KS66}, \cite{Ku09}, \cite{KK04}, \cite{LT77}, \cite{MM06}, \cite{Ma04}, \cite{Ne83}, \cite{Pa06}, \cite{Pa87}, \cite{Pa02}, \cite{Pa08}, \cite{Po01}, \cite{Po03}, \cite{Po04}, \cite{Po08}, \cite{PR09}, \cite{Ry07}, \cite{Ry09}, \cite{Ry10}, \cite{Sa65}, \cite{St70a}, \cite{TS77}, and the references cited therein. We start by explicitly computing operators of the type $\gamma_{\theta_0^{\prime},\theta_R^{\prime}}(\ensuremath{H_{\te_0,\te_R}} - z I)^{-1}$ which play a role at various places in this manuscript (cf.\ Theorem \ref{t3.5}, Lemma \ref{l6.2}, and Theorem \ref{t6.3}): \begin{lemma} \label{l6.1} Assume that $\theta_0, \theta_R, \theta_0', \theta_R' \in S_{2 \pi}$, let $\ensuremath{H_{\te_0,\te_R}}$ be defined as in \eqref{2.2a}, and suppose that $z\in{\mathbb{C}}\backslash\sigma(\ensuremath{H_{\te_0,\te_R}})$. Then, assuming $f \in L^2((0,R); dx)$, and writing \begin{align} \gamma_{\theta_0^{\prime},\theta_R^{\prime}}(\ensuremath{H_{\te_0,\te_R}} - z I)^{-1} f = \begin{bmatrix} \big(\gamma_{\theta_0^{\prime},\theta_R^{\prime}}(\ensuremath{H_{\te_0,\te_R}} - z I)^{-1}\big)_1f \\[2mm] \big(\gamma_{\theta_0^{\prime},\theta_R^{\prime}}(\ensuremath{H_{\te_0,\te_R}} - z I)^{-1}\big)_2 f \end{bmatrix} \in {\mathbb{C}}^2, \label{6.1} \end{align} one has \begin{align} \big(\gamma_{\theta_0^{\prime},\theta_R^{\prime}}(\ensuremath{H_{\te_0,\te_R}} - z I)^{-1}\big)_1f &= \frac{\sin(\theta_0'-\theta_0)}{W(u_{+,\theta_R}(z,\cdot), u_{-,\theta_0}(z,\cdot))} \big(\overline{u_{+,\theta_R}(z,\cdot)}, f)_{L^2((0,R); dx)} \notag \\ & \quad \times \begin{cases} - \frac{u_{-,\theta_0}(z,0)}{\sin(\theta_0)}, & \theta_0 \in S_{2 \pi}\backslash\{0,\pi\}, \\ \frac{u'_{-,\theta_0}(z,0)}{\cos(\theta_0)}, & \theta_0 \in S_{2 \pi}\backslash\{\pi/2,3\pi/2\}, \end{cases} \label{6.2} \\ \big(\gamma_{\theta_0^{\prime},\theta_R^{\prime}}(\ensuremath{H_{\te_0,\te_R}} - z I)^{-1}\big)_2 f &= \frac{\sin(\theta_R'-\theta_R)}{W(u_{+,\theta_R}(z,\cdot), u_{-,\theta_0}(z,\cdot))} \big(\overline{u_{-,\theta_0}(z,\cdot)}, f)_{L^2((0,R); dx)} \notag \\ & \quad \times \begin{cases} - \frac{u_{+,\theta_R}(z,R)}{\sin(\theta_R)}, & \theta_R \in S_{2 \pi}\backslash\{0,\pi\}, \\ \frac{u'_{+,\theta_R}(z,R)}{\cos(\theta_R)}, & \theta_R \in S_{2 \pi}\backslash\{\pi/2,3\pi/2\}, \end{cases} \label{6.3} \end{align} in particular, \begin{align} & |\big(\gamma_{\theta_0^{\prime},\theta_R^{\prime}}(\ensuremath{H_{\te_0,\te_R}} - z I)^{-1}\big)_1f| \underset{\theta_0'\to \theta_0}{=} O(\theta_0' - \theta_0) C_1(z) \|f\|_{L^2((0,R); dx)}, \label{6.4} \\ & |\big(\gamma_{\theta_0^{\prime},\theta_R^{\prime}}(\ensuremath{H_{\te_0,\te_R}} - z I)^{-1}\big)_2f| \underset{\theta_R'\to \theta_R}{=} O(\theta_R' - \theta_R) C_2(z) \|f\|_{L^2((0,R); dx)}, \label{6.5} \end{align} for some constants $C_j(z) > 0$, $j=1,2$, and hence \begin{align} & \gamma_{\theta_0,\theta_R}(\ensuremath{H_{\te_0,\te_R}} - z I)^{-1} = 0 \, \text{ in ${\mathcal B}\big(L^2((0,R); dx), {\mathbb{C}}^2\big)$}, \label{6.5a} \\ & \big(\gamma_{\theta_0,\theta_R}(\ensuremath{H_{\te_0,\te_R}} - z I)^{-1}\big)_k = 0 \, \text{ in ${\mathcal B}\big(L^2((0,R); dx), {\mathbb{C}}\big)$}, \; k=1,2. \label{6.5b} \end{align} \end{lemma} \begin{proof} Employing \eqref{3.32} and \eqref{3.33} one obtains \begin{align} \gamma_{\theta_0^{\prime},\theta_R^{\prime}}(\ensuremath{H_{\te_0,\te_R}} - z I)^{-1} f & = \gamma_{\theta_0^{\prime},\theta_R^{\prime}} \bigg(\int_0^R dx' \, G_{\theta_0,\theta_R}(z,\cdot,x') f(x')\bigg) \notag \\ & = \begin{bmatrix} \bigg(\gamma_{\theta_0^{\prime},\theta_R^{\prime}} \bigg(\int_0^R dx' \, G_{\theta_0,\theta_R}(z,\cdot,x') f(x')\bigg)\bigg)_1 \\[4mm] \bigg(\gamma_{\theta_0^{\prime},\theta_R^{\prime}} \bigg(\int_0^R dx' \, G_{\theta_0,\theta_R}(z,\cdot,x') f(x')\bigg)\bigg)_2 \end{bmatrix} \label{6.6} \end{align} and hence (with $W(z) = W(u_{+,\theta_R}(z,\cdot), u_{-,\theta_0}(z,\cdot))$) \begin{align} & \bigg(\gamma_{\theta_0^{\prime},\theta_R^{\prime}} \bigg(\int_0^R dx' \, G_{\theta_0,\theta_R}(z,\cdot,x') f(x')\bigg)\bigg)_1 \notag \\ & \quad = \frac{1}{W(z)} \bigg\{\cos(\theta_0') \bigg(\int_0^R dx' \, G_{\theta_0,\theta_R}(z,0,x') f(x')\bigg) \notag \\ & \hspace*{1.95cm} + \sin(\theta_0') \bigg(\int_0^R dx' \, \bigg(\frac{\partial}{\partial x} G_{\theta_0,\theta_R}(z,x,x')\bigg|_{x<x'}\bigg)\bigg|_{x=0} f(x')\bigg)\bigg\} \notag \\ & \quad = \frac{1}{W(z)} \bigg\{\big[\cos(\theta_0') u_{-,\theta_0}(z,0) + \sin(\theta_0') u_{+,\theta_0}'(z,0)\big] \int_0^R dx' \, u_{+,\theta_R}(z,x') f(x')\bigg\} \notag \\ & \quad = \frac{1}{W(z)} \big(\overline{u_{+,\theta_R}(z,\cdot)}, f)_{L^2((0,R); dx)} \notag \\ & \qquad \, \times \begin{cases} [\cos(\theta_0') - \sin(\theta_0') \cot(\theta_0)] u_{-,\theta_0}(z,0), & \theta_0 \in S_{2 \pi}\backslash\{0,\pi\}, \\ [- \cos(\theta_0') \tan(\theta_0) + \sin(\theta_0')] u_{-,\theta_0}'(z,0), & \theta_0 \in S_{2 \pi}\backslash\{\pi/2,3\pi/2\}. \end{cases} \label{6.7} \end{align} Equation \eqref{6.7} is easily seen to be equivalent to \eqref{6.2}. Equation \eqref{6.3} is derived analogously. \end{proof} Introducing the orthogonal projections in ${\mathbb{C}}^2$, \begin{equation} P_1 = \begin{bmatrix} 1 & 0 \\ 0 & 0 \end{bmatrix}, \quad P_2 = \begin{bmatrix} 0 & 0 \\ 0 & 1 \end{bmatrix}, \label{6.8} \end{equation} one obtains the following result, patterned after \cite[Lemma\ 6]{Na01} in the context of Schr\"odinger operators with Dirichlet and Neumann boundary conditions on a cube in ${\mathbb{R}}^n$: \begin{lemma} \label{l6.2} Assume that $\theta_0, \theta_R, \theta_0', \theta_R' \in S_{2 \pi}$, let $\ensuremath{H_{\te_0,\te_R}}$ and $H_{\theta_0',\theta_R'}$ be defined as in \eqref{2.2a}, and suppose that $z\in{\mathbb{C}}\big\backslash\big(\sigma(\ensuremath{H_{\te_0,\te_R}})\cup \sigma(H_{\theta_0',\theta_R'})\big)$. Then \begin{align} & (H_{\theta_0',\theta_R'} - z I)^{-1} = (\ensuremath{H_{\te_0,\te_R}} - z I)^{-1} \notag \\ & \quad + \big[\gamma_{\overline{\theta_0'},\overline{\theta_R'}} ((\ensuremath{H_{\te_0,\te_R}})^* - {\overline z} I)^{-1}\big]^* S_{\theta_0'-\theta_0, \theta_R'-\theta_R}^{-1} \big[\gamma_{\theta_0,\theta_R} (H_{\theta_0',\theta_R'} - z I)^{-1}\big], \label{6.9} \\ & \hspace*{7.95cm} \theta_0' \neq \theta_0, \; \theta_R' \neq \theta_R, \notag \\ & (H_{\theta_0,\theta_R'} - z I)^{-1} = (\ensuremath{H_{\te_0,\te_R}} - z I)^{-1} \notag \\ & \quad + \big[\gamma_{\overline{\theta_0},\overline{\theta_R'}} ((\ensuremath{H_{\te_0,\te_R}})^* - {\overline z} I)^{-1}\big]^* [\sin(\theta_R' - \theta_R)]^{-1}P_2 \big[\gamma_{\theta_0,\theta_R} (H_{\theta_0,\theta_R'} - z I)^{-1}\big], \notag \\ & \hspace*{9.5cm} \theta_R' \neq \theta_R, \label{6.10} \\ & (H_{\theta_0',\theta_R} - z I)^{-1} = (\ensuremath{H_{\te_0,\te_R}} - z I)^{-1} \notag \\ & \quad + \big[\gamma_{\overline{\theta_0'},\overline{\theta_R}} ((\ensuremath{H_{\te_0,\te_R}})^* - {\overline z} I)^{-1}\big]^* [\sin(\theta_0' - \theta_0)]^{-1}P_1 \big[\gamma_{\theta_0,\theta_R} (H_{\theta_0',\theta_R} - z I)^{-1}\big], \notag \\ & \hspace*{9.6cm} \theta_0' \neq \theta_0. \label{6.11} \end{align} \end{lemma} \begin{proof} We first consider the case \begin{equation} \theta_0, \theta_R, \theta_0', \theta_R' \in S_{2 \pi}\backslash\{0,\pi\}, \quad \theta_0' \neq \theta_0, \quad \theta_R' \neq \theta_R, \label{6.12} \end{equation} which illustrates the principal idea of the proof. To get started, we pick $f, g \in L^2((0,R); dx)$ and introduce \begin{align} \begin{split} \phi &= ((\ensuremath{H_{\te_0,\te_R}})^* - \overline z I)^{-1} f \in \text{\rm{dom}}((\ensuremath{H_{\te_0,\te_R}})^*), \\ \psi &= (H_{\theta_0',\theta_R'} - z I)^{-1} g \in \text{\rm{dom}}(H_{\theta_0',\theta_R'}). \label{6.13} \end{split} \end{align} Then one computes \begin{align} & ((\ensuremath{H_{\te_0,\te_R}})^* - \overline z I) \phi, \psi)_{L^2((0,R); dx)} - (\phi, (H_{\theta_0',\theta_R'} - z I) \psi)_{L^2((0,R); dx)} \notag \\ & \quad = - \int_0^R dx \, \overline{\phi''(x)} \psi(x) + \int_0^R dx \, \overline{\phi(x)} \psi''(x) \notag \\ & \quad = - \overline{\phi'(R)} \psi(R) + \overline{\phi'(0)} \psi(0) + \overline{\phi(R)} \psi'(R) - \overline{\phi(0)} \psi'(0) \notag \\ & \quad = [\cot(\theta_R')-\cot(\theta_R)] \overline{\phi(R)} \psi(R) + [\cot(\theta_0')-\cot(\theta_0)] \overline{\phi(0)} \psi(0) \notag \\ & \quad = - \frac{\sin(\theta_R'-\theta_R)}{\sin(\theta_R')\sin(\theta_R)} \overline{\phi(R)} \psi(R) - \frac{\sin(\theta_0'-\theta_0)}{\sin(\theta_0')\sin(\theta_0)} \overline{\phi(0)} \psi(0), \label{6.14} \end{align} using the fact that \eqref{6.13} implies \begin{align} \begin{split} \cos(\overline \theta_0) \phi(0) + \sin(\overline \theta_0) \phi'(0) &= 0, \\ \cos(\overline \theta_R) \phi(R) - \sin(\overline \theta_R) \phi'(R) &= 0, \\ \cos(\theta_0') \psi(0) + \sin(\theta_0') \psi'(0) &= 0, \label{6.15} \\ \cos(\theta_R') \psi(R) - \sin(\theta_R') \psi'(R) &= 0. \end{split} \end{align} Using \eqref{6.15} once again, one also computes \begin{align} (\gamma_{\overline{\theta_0'},\overline{\theta_R'}} \phi, \gamma_{\theta_0,\theta_R} \psi)_{{\mathbb{C}}^2} = - \frac{\sin^2(\theta_R'-\theta_R)}{\sin(\theta_R')\sin(\theta_R)} \overline{\phi(R)} \psi(R) - \frac{\sin^2(\theta_0'-\theta_0)}{\sin(\theta_0')\sin(\theta_0)} \overline{\phi(0)} \psi(0). \label{6.16} \end{align} A comparison of \eqref{6.14} and \eqref{6.16} then yields \begin{align} & ((\ensuremath{H_{\te_0,\te_R}})^* - \overline z I) \phi, \psi)_{L^2((0,R); dx)} - (\phi, (H_{\theta_0',\theta_R'} - z I) \psi)_{L^2((0,R); dx)} \notag \\ & \quad = \big(\gamma_{\overline{\theta_0'},\overline{\theta_R'}} \phi, S_{\theta_0'-\theta_0,\theta_R'-\theta_R}^{-1} \gamma_{\theta_0,\theta_R} \psi\big)_{{\mathbb{C}}^2}, \label{6.17} \end{align} or equivalently, \begin{align} & \big(f,(H_{\theta_0',\theta_R'} - z I)^{-1} g\big)_{L^2((0,R); dx)} = \big(f, (\ensuremath{H_{\te_0,\te_R}} - z I)^{-1} g\big)_{L^2((0,R); dx)} \notag \\ & \quad + \big(f, \big[\gamma_{\overline{\theta_0'},\overline{\theta_R'}} ((\ensuremath{H_{\te_0,\te_R}})^* - {\overline z} I)^{-1}\big]^* S_{\theta_0'-\theta_0, \theta_R'-\theta_R}^{-1} \label{6.18} \\ & \qquad \times \big[\gamma_{\theta_0,\theta_R} (H_{\theta_0',\theta_R'} - z I)^{-1}\big] g)_{L^2((0,R); dx)}, \notag \end{align} and hence \eqref{6.9} since $f, g \in L^2((0,R); dx)$ are arbitrary, under the additional assumptions in \eqref{6.12}. Employing Lemma \ref{l6.1} (and particularly, \eqref{6.4}--\eqref{6.5b}) then shows that \eqref{6.9} is continuous in $\theta_0, \theta_R, \theta_0', \theta_R' \in S_{2 \pi}$ with respect to the norm in ${\mathcal B}\big(L^2((0,R); dx)\big)$, removing the restrictions $\theta_0, \theta_R, \theta_0', \theta_R' \in S_{2 \pi}\backslash\{0,\pi\}$ in \eqref{6.18}, and thus proving \eqref{6.9}. Analogous considerations imply \eqref{6.10} and \eqref{6.11}. \end{proof} The principal result of this section, Krein's formula for the difference of resolvents of $H_{\theta_0',\theta_R'}$ and $\ensuremath{H_{\te_0,\te_R}}$, then reads as follows: \begin{theorem} \label{t6.3} Assume that $\theta_0, \theta_R, \theta_0', \theta_R' \in S_{2 \pi}$, let $\ensuremath{H_{\te_0,\te_R}}$ and $H_{\theta_0',\theta_R'}$ be defined as in \eqref{2.2a}, and suppose that $z\in{\mathbb{C}}\big\backslash\big(\sigma(\ensuremath{H_{\te_0,\te_R}})\cup \sigma(H_{\theta_0',\theta_R'})\big)$. Then, with $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z)$ introduced in \eqref{2.6}, \begin{align} & (H_{\theta_0',\theta_R'} -z I)^{-1} = (\ensuremath{H_{\te_0,\te_R}} -z I)^{-1} \notag \\ & \quad - \big[\gamma_{\overline{\theta_0^{\prime}},\overline{\theta_R^{\prime}}} ((\ensuremath{H_{\te_0,\te_R}})^* - {\overline z} I)^{-1}\big]^* S_{\theta'_0-\theta'_0,\theta'_R-\theta_R}^{-1} \label{6.21} \\ & \qquad \times \Big[\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z)\Big]^{-1} \big[\gamma_{\theta_0^{\prime},\theta_R^{\prime}}(\ensuremath{H_{\te_0,\te_R}} - z I)^{-1}\big], \quad \theta_0 \neq \theta_0', \; \theta_R \neq \theta_R'. \notag \\ & (H_{\theta_0,\theta_R'} -z I)^{-1} = (\ensuremath{H_{\te_0,\te_R}} -z I)^{-1} \notag \\ & \quad - \big[\gamma_{\overline{\theta_0},\overline{\theta_R^{\prime}}} ((\ensuremath{H_{\te_0,\te_R}})^* - {\overline z} I)^{-1}\big]^* [\sin(\theta'_R-\theta_R)]^{-1} P_2 \label{6.22} \\ & \qquad \times \Big[\Lambda_{\theta_0,\theta_R}^{\theta_0,\theta_R'} (z)\Big]^{-1} P_2 \big[\gamma_{\theta_0,\theta_R^{\prime}}(\ensuremath{H_{\te_0,\te_R}} - z I)^{-1}\big], \quad \theta_R \neq \theta_R', \notag \\ & (H_{\theta_0',\theta_R} -z I)^{-1} = (\ensuremath{H_{\te_0,\te_R}} -z I)^{-1} \notag \\ & \quad - \big[\gamma_{\overline{\theta_0'},\overline{\theta_R}} ((\ensuremath{H_{\te_0,\te_R}})^* - {\overline z} I)^{-1}\big]^* [\sin(\theta_0'-\theta_0)]^{-1} P_1 \label{6.23} \\ & \qquad \times \Big[\Lambda_{\theta_0,\theta_R}^{\theta_0',\theta_R} (z)\Big]^{-1} P_1 \big[\gamma_{\theta_0',\theta_R}(\ensuremath{H_{\te_0,\te_R}} - z I)^{-1}\big], \quad \theta_0 \neq \theta_0'. \notag \end{align} \end{theorem} \begin{proof} Taking adjoints on both sides of \eqref{6.9}, and subsequently replacing $z$ by $\overline z$, $\theta_0, \theta_R$ by $\overline \theta_0, \overline \theta_R$, and $V$ by $\overline V$, then yields \begin{align} & (H_{\theta_0',\theta_R'} - z I)^{-1} = (\ensuremath{H_{\te_0,\te_R}} - z I)^{-1} \notag \\ & \quad + \big[\gamma_{\overline{\theta_0},\overline{\theta_R}} ((H_{\theta_0',\theta_R'})^* - {\overline z} I)^{-1}\big]^* S_{\theta_0'-\theta_0, \theta_R'-\theta_R}^{-1} \big[\gamma_{\theta_0',\theta_R'} (\ensuremath{H_{\te_0,\te_R}} - z I)^{-1}\big], \label{6.24} \\ & \hspace*{7.95cm} \theta_0' \neq \theta_0, \; \theta_R' \neq \theta_R. \notag \end{align} Applying $\gamma_{\theta_0,\theta_R}$ to both sides of \eqref{6.24}, and using the fact that $\gamma_{\theta_0,\theta_R} (\ensuremath{H_{\te_0,\te_R}} - z I)^{-1} = 0$ by \eqref{6.5a}, one obtains \begin{align} & \gamma_{\theta_0,\theta_R} (H_{\theta_0',\theta_R'} - z I)^{-1} = \gamma_{\theta_0,\theta_R} \big[\gamma_{\overline{\theta_0},\overline{\theta_R}} ((H_{\theta_0',\theta_R'})^* - {\overline z} I)^{-1}\big]^* \notag \\ & \quad \times S_{\theta_0'-\theta_0, \theta_R'-\theta_R}^{-1} \big[\gamma_{\theta_0',\theta_R'} (\ensuremath{H_{\te_0,\te_R}} - z I)^{-1}\big], \quad \theta_0' \neq \theta_0, \; \theta_R' \neq \theta_R. \label{6.25} \end{align} An insertion of \eqref{6.25} into \eqref{6.9} implies \begin{align} & (H_{\theta_0',\theta_R'} - z I)^{-1} = (\ensuremath{H_{\te_0,\te_R}} - z I)^{-1} \notag \\ & \quad + \big[\gamma_{\overline{\theta_0'},\overline{\theta_R'}} ((\ensuremath{H_{\te_0,\te_R}})^* - {\overline z} I)^{-1}\big]^* S_{\theta_0'-\theta_0, \theta_R'-\theta_R}^{-1} \gamma_{\theta_0,\theta_R} \big[\gamma_{\overline{\theta_0},\overline{\theta_R}} ((H_{\theta_0',\theta_R'})^* - {\overline z} I)^{-1}\big]^* \notag \\ & \qquad \times S_{\theta_0'-\theta_0, \theta_R'-\theta_R}^{-1} \big[\gamma_{\theta_0',\theta_R'} (\ensuremath{H_{\te_0,\te_R}} - z I)^{-1}\big], \quad \theta_0' \neq \theta_0, \; \theta_R' \neq \theta_R. \label{6.26} \end{align} Using \eqref{2.48} and \eqref{3.33aa} one obtains \begin{equation} \gamma_{\theta_0,\theta_R} \big[\gamma_{\overline{\theta_0},\overline{\theta_R}} ((H_{\theta_0',\theta_R'})^*- {\overline z} I)^{-1}\big]^* = - \Big[\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z)\Big]^{-1} S_{\theta_0'-\theta_0,\theta_R'-\theta_R}, \quad z \in {\mathbb{C}}_+, \label{6.27} \end{equation} and inserting \eqref{6.27} into \eqref{6.26} yields \eqref{6.21}. Since equations \eqref{6.22} and \eqref{6.23} are proved similarly, we just briefly sketch the proof of \eqref{6.22}: First, in analogy to \eqref{6.24}, one derives from \eqref{6.10} that \begin{align} & (H_{\theta_0,\theta_R'} - z I)^{-1} = (\ensuremath{H_{\te_0,\te_R}} - z I)^{-1} \notag \\ & \quad + \big[\gamma_{\overline{\theta_0},\overline{\theta_R}} ((H_{\theta_0,\theta_R'})^* - {\overline z} I)^{-1}\big]^* [\sin(\theta_R' - \theta_R)]^{-1}P_2 \big[\gamma_{\theta_0,\theta_R'} (H_{\theta_0,\theta_R} - z I)^{-1}\big], \notag \\ & \hspace*{9.5cm} \theta_R' \neq \theta_R. \label{6.28} \end{align} Applying $\gamma_{\theta_0,\theta_R}$ to both sides of \eqref{6.28}, and using again the fact that $\gamma_{\theta_0,\theta_R} (\ensuremath{H_{\te_0,\te_R}} - z I)^{-1} = 0$, one obtains \begin{align} & \gamma_{\theta_0,\theta_R} (H_{\theta_0,\theta_R'} - z I)^{-1} = \gamma_{\theta_0,\theta_R} \big[\gamma_{\overline{\theta_0},\overline{\theta_R}} ((H_{\theta_0,\theta_R'})^* - {\overline z} I)^{-1}\big]^* \notag \\ & \quad \times [\sin(\theta_R'-\theta_R)]^{-1} P_2 \big[\gamma_{\theta_0,\theta_R'} (\ensuremath{H_{\te_0,\te_R}} - z I)^{-1}\big], \quad \theta_R' \neq \theta_R. \label{6.29} \end{align} Finally, an insertion of \eqref{6.29} into the right-hand side of \eqref{6.10}, and using \eqref{6.27} in the special case $\theta_0'=\theta_0$, that is, \begin{align} & \gamma_{\theta_0,\theta_R} \big[\gamma_{\overline{\theta_0},\overline{\theta_R}} ((H_{\theta_0,\theta_R'})^*- {\overline z} I)^{-1}\big]^* = - \Big[\Lambda_{\theta_0,\theta_R}^{\theta_0,\theta_R'} (z)\Big]^{-1} S_{0,\theta_R'-\theta_R} \notag \\ & \quad = - \Big[\Lambda_{\theta_0,\theta_R}^{\theta_0,\theta_R'} (z)\Big]^{-1} [\sin(\theta_R'-\theta_R)]^{-1} P_2, \quad z \in {\mathbb{C}}_+, \end{align} yields \eqref{6.22}. \end{proof} \section{A Brief Outlook} \label{s7} In this section we provide a brief comparison between $\ensuremath{\La_{\te_0,\te_R}} (z)$ and the $2 \times 2$ matrix-valued Weyl--Titchmarsh function associated with $\ensuremath{H_{\te_0,\te_R}}$ in the self-adjoint context, that is, under the assumptions $\theta_0, \theta_R \in [0, 2\pi)$ and $V$ is real-valued in addition to \eqref{2.2aa}. While both $2 \times 2$ matrices are matrix-valued Herglotz functions, they are quite different as this brief section will show. A more detailed discussion of the interrelations between these two matrices is beyond the scope of this paper and will be taken up elsewhere. To exhibit some of the differences between $\ensuremath{\La_{\te_0,\te_R}} (z)$ and various versions of the Weyl--Titchmarsh matrix we will link the entries in both matrices to the Green's function $G_{\theta_0,\theta_R} (z,x,x')$ of $\ensuremath{H_{\te_0,\te_R}}$ and provide some formulas which may well be of independent interest. We start by linking $\ensuremath{\La_{\te_0,\te_R}} (z)$ and $G_{\theta_0,\theta_R} (z,x,x')$ and list a variety of pertinent formulas (choosing $z\in{\mathbb{C}}\backslash{\mathbb{R}}$ for notational simplicity): \begin{align} \ensuremath{\La_{\te_0,\te_R}} (z)_{1,1} &= m_{+,\theta_0}(z,\theta_R) \notag \\ &= \dfrac{-\sin(\theta_0) + \cos(\theta_0)m_{+,0}(z,\theta_R)} {\cos(\theta_0) + \sin(\theta_0)m_{+,0}(z,\theta_R)} \notag \\[1mm] &= \dfrac{-\sin(\theta_0) + \cos(\theta_0)u_{+,\theta_R}'(z,0)}{\cos(\theta_0) + \sin(\theta_0)u_{+,\theta_R}'(z,0)} \notag \\[1mm] &= \frac{1}{\sin^2(\theta_0)} \big[G_{\theta_0,\theta_R} (z,0,0) + \sin(\theta_0)\cos(\theta_0)\big], \label{7.1}\\ &\hspace*{1.95cm} \theta_0 \in [0, 2\pi)\backslash\{0,\pi\}, \, \theta_R \in [0, 2\pi), \notag \\ \Lambda_{0,0} (z)_{1,1} & = m_{+,0}(z,0) \notag \\ &= \lim_{0<x<x', \, x'\downarrow 0} \partial_x \partial_{x'} G_{0,0} (z,x,x'). \label{7.2} \end{align} Here we used \eqref{3.24} and (cf.\ also \eqref{3.11b}) \begin{align} \begin{split} G_{\theta_0,\theta_R} (z,0,0) &= \frac{- \sin(\theta_0)}{\cos(\theta_0) + \sin(\theta_0) m_{+,0}(z,\theta_R)} \\[1mm] &= \sin(\theta_0) \big[- \cos(\theta_0) + \sin(\theta_0) m_{+,\theta_0}(z,\theta_R)\big]. \label{7.3} \end{split} \end{align} In the same manner one obtains \begin{align} \ensuremath{\La_{\te_0,\te_R}} (z)_{2,2} &= - m_{-,\theta_R}(z,\theta_0) \notag \\ &= - \dfrac{\sin(\theta_R) + \cos(\theta_R)m_{-,0}(z,\theta_0)} {\cos(\theta_R) - \sin(\theta_R)m_{-,0}(z,\theta_0)} \notag \\[1mm] &= - \dfrac{\sin(\theta_R) + \cos(\theta_R)u_{-,\theta_0}'(z,R)}{\cos(\theta_R) - \sin(\theta_R)u_{-,\theta_0}'(z,R)} \notag \\[1mm] &= \frac{1}{\sin^2(\theta_R)} \big[G_{\theta_0,\theta_R} (z,R,R) + \sin(\theta_R)\cos(\theta_R)\big], \label{7.4}\\ &\hspace*{2.35cm} \theta_0 \in [0, 2\pi), \, \theta_R \in [0, 2\pi)\backslash\{0,\pi\}, \notag \\ \Lambda_{0,0} (z)_{2,2} &= - m_{-,0}(z,0) \notag \\ &= \lim_{0<x<x', \, x\uparrow R} \partial_x \partial_{x'} G_{0,0} (z,x,x'). \label{7.5} \end{align} Here we used \eqref{3.28} and (cf.\ also \eqref{3.25}) \begin{align} \begin{split} G_{\theta_0,\theta_R} (z,R,R) &= \frac{- \sin(\theta_R)}{\cos(\theta_R) - \sin(\theta_R) m_{-,0}(z,\theta_0)} \\[1mm] &= \sin(\theta_R) \big[- \cos(\theta_R) - \sin(\theta_R) m_{-,\theta_R}(z,\theta_0)\big]. \label{7.6} \end{split} \end{align} Similarly, the off-diagonal terms of $\ensuremath{\La_{\te_0,\te_R}} (z)$ in \eqref{3.11a} can be written as \begin{align} & \ensuremath{\La_{\te_0,\te_R}} (z)_{1,2} = \ensuremath{\La_{\te_0,\te_R}} (z)_{2,1} \notag \\ & \quad = \frac{1}{\sin(\theta_R)} G_{\theta_0,\theta_R} (z,R,R) \begin{cases} \frac{- u_{-,\theta_0}'(z,0)}{\cos(\theta_0)}, & \theta_0 \in [0, 2\pi)\backslash\{\pi/2, 3\pi/2\}, \\ \frac{u_{-,\theta_0}(z,0)}{\sin(\theta_0)}, & \theta_0 \in [0, 2\pi)\backslash\{0, \pi\},\end{cases} \label{7.7} \\ & \hspace{6.65cm} \theta_R \in [0, 2\pi)\backslash\{0,\pi\}, \notag \\ & \quad = \frac{1}{\sin(\theta_0)} G_{\theta_0,\theta_R} (z,0,0) \begin{cases} \frac{u_{+,\theta_R}'(z,R)}{\cos(\theta_R)}, & \theta_R \in [0, 2\pi)\backslash\{\pi/2, 3\pi/2\}, \\ \frac{u_{+,\theta_R}(z,R)}{\sin(\theta_R)}, & \theta_R \in [0, 2\pi)\backslash\{0, \pi\},\end{cases} \label{7.8} \\ & \hspace{6.45cm} \theta_0 \in [0, 2\pi)\backslash\{0,\pi\}. \notag \end{align} Next we turn to (variants of) the $2 \times 2$ Weyl--Titchmarsh matrix corresponding to $\ensuremath{H_{\te_0,\te_R}}$ with respect to an interior reference point $x_0 \in (0,R)$. We start by introducing (again, choosing $z\in{\mathbb{C}}\backslash{\mathbb{R}}$ for notational simplicity) \begin{align} &m_{+,0} (z, x_0,\theta_R) = \frac{u_{+,\theta_R}' (z,x_0)}{u_{+,\theta_R} (z,x_0)}, \quad x_0 \in (0,R), \label{7.9} \\ &m_{-,0} (z, x_0,\theta_0) = \frac{u_{-,\theta_0}'(z,x_0)}{u_{-,\theta_0} (z,x_0)}, \quad x_0 \in (0,R), \label{7.10} \end{align} and more generally, \begin{align} m_{+,\alpha}(z,x_0,\theta_R) &=\frac{-\sin(\alpha) + \cos(\alpha) m_{+,0}(z,x_0,\theta_R)} {\cos(\alpha) +\sin(\alpha) m_{+,0}(z,x_0,\theta_R)}, \quad \alpha \in[0,\pi), \label{7.11} \\ m_{-,\alpha}(z,x_0,\theta_0) &=\frac{-\sin(\alpha) + \cos(\alpha) m_{-,0}(z,x_0,\theta_0)} {\cos(\alpha) +\sin(\alpha) m_{-,0}(z,x_0,\theta_0)}, \quad \alpha \in[0,\pi). \label{7.12} \end{align} We note that $m_{+,\alpha}(\cdot,x_0,\theta_R)$ and $- m_{-,\alpha}(\cdot,x_0,\theta_0)$ are known to be Herglotz functions (cf., e.g., \cite[Sect.\ 9.5]{CL85}), \cite[Sect.\ II.8]{LS75}, \cite[Sect.\ 6.5]{Pe88}, \cite[Ch.\ III]{Ti62}). Associated with \eqref{7.9}--\eqref{7.12} one then defines the $2 \times 2$ Weyl--Titchmarsh matrices (the fundamental ingredient for inverse spectral theory for operators of the type $\ensuremath{H_{\te_0,\te_R}}$) by, \begin{align} &M_{0}(z,x_0,\theta_0,\theta_R) = \big[M_{0, j,k}(z,x_0,\theta_0,\theta_R)\big]_{j,k=1,2}\notag \\ &= \big[m_{-,0}(z,x_0,\theta_0)-m_{+,}(z,x_0,\theta_R)\big]^{-1} \label{7.13} \\ & \quad \times\begin{bmatrix} 1 & \f12[m_{-,0}(z,x_0,\theta_0) +m_{+,0}(z,x_0,\theta_R)] \\ \f12[m_{-,0}(z,x_0,\theta_0)+m_{+,0}(z,x_0,\theta_R)] & m_{-,0}(z,x_0,\theta_0)m_{+,0}(z,x_0,\theta_R) \end{bmatrix}, \notag \end{align} and more generally, by \begin{align} &M_{\alpha}(z,x_0,\theta_0,\theta_R) = \big[M_{\alpha, j,k}(z,x_0,\theta_0,\theta_R)\big]_{j,k=1,2}\notag \\ &= \big[m_{-,\alpha}(z,x_0,\theta_0)-m_{+,\alpha}(z,x_0,\theta_R)\big]^{-1} \label{7.14} \\ & \quad \times\begin{bmatrix} 1 & \f12[m_{-,\alpha}(z,x_0,\theta_0) +m_{+,\alpha}(z,x_0,\theta_R)] \\ \f12[m_{-,\alpha}(z,x_0,\theta_0)+m_{+,\alpha}(z,x_0,\theta_R)] & m_{-,\alpha}(z,x_0,\theta_0)m_{+,\alpha}(z,x_0,\theta_R) \end{bmatrix}, \notag \\ & \hspace*{11cm} \alpha \in[0,\pi). \notag \end{align} By inspection, \begin{align} \det (M_{\alpha}(z,x_0,\theta_0,\theta_R))&= -1/4, \quad \alpha \in[0,\pi), \; z\in{\mathbb{C}}\setminus{\mathbb{R}}, \label{7.15} \\ \Im(M_{\alpha}(z,x_0,\theta_0,\theta_R)) &> 0, \quad \alpha \in[0,\pi), \; z\in{\mathbb{C}}_+, \label{7.16} \end{align} and hence $M_{\alpha} (\cdot,x_0,\theta_0,\theta_R)$ is a $2\times 2$ matrix-valued Herglotz function implying that $M_{\alpha, j,j}(\cdot,x_0,\theta_0,\theta_R)$ are Herglotz functions for $j=1,2$. In particular, the matrices $M_{\alpha} (z,x_0,\theta_0,\theta_R)$ can be shown to have Herglotz representations of the type \eqref{5.1}. For the connection of $M_{0} (z,x_0,\theta_0,\theta_R)$ and $M_{\alpha} (z,x_0,\theta_0,\theta_R)$ with the Green's function $G_{\theta_0,\theta_R} (z,x,x')$ of $\ensuremath{H_{\te_0,\te_R}}$ we first introduce a bit of notation: \begin{align} \partial_{1} G_{\theta_0,\theta_R}(z,x_0, x') & = \partial_{x_1} G_{\theta_0,\theta_R}(z,x_1,x')\big|_{x_1=x_0}, \notag \\ \partial_{2} G_{\theta_0,\theta_R}(z,x,x_0) &= \partial_{x_2} G_{\theta_0,\theta_R} (z,x,x_2)\big|_{x_2 = x_0}, \label{7.17} \\ \partial_{1} \partial_{2} G_{\theta_0,\theta_R} (z,x_0, x_0) & = \partial_{x_1} \partial_{x_2} G_{\theta_0,\theta_R} (z,x_1,x_2)\big|_{x_1=x_0, x_2=x_0},\; \mbox{ etc.} \notag \end{align} The expressions \eqref{7.13} and \eqref{7.14} for $M_0 (z,x_0,\theta_0,\theta_R)$ and $M_\alpha(z,x_0,\theta_0,\theta_R)$ then can be rewritten as follows: \begin{align} M_{0,1,1}(z,x_0,\theta_0,\theta_R) &= G_{\theta_0,\theta_R}(z,x_0,x_0). \label{7.18} \\ M_{0,1,2}(z,x_0,\theta_0,\theta_R) &= M_{0,2,1}(z,x_0,\theta_0,\theta_R) \notag \\ & = (1/2) (\partial_1 + \partial_2) G_{\theta_0,\theta_R}(z,x_0\pm 0,x_0\mp 0), \label{7.19} \\ M_{0,2,2}(z,x_0,\theta_0,\theta_R) &= \partial_1 \partial_2 G_{\theta_0,\theta_R}(z,x_0,x_0), \label{7.20} \end{align} and \begin{align} &M_{\alpha,1,1}(z,x_0,\theta_0,\theta_R) \notag \\ & \quad =\big(\cos(\alpha) +\sin(\alpha)\partial_1\big)\big(\cos(\alpha)+ \sin(\alpha)\partial_2\big)G_{\theta_0,\theta_R}(z,x_0,x_0). \label{7.21} \\ &M_{\alpha,1,2}(z,x_0,\theta_0,\theta_R)=M_{\alpha,2,1}(z,x_0,\theta_0,\theta_R) \notag \\ & \quad =(1/2)\big((\cos(\alpha)+\sin(\alpha)\partial_1) (-\sin(\alpha)+\cos(\alpha)\partial_2) \notag \\ &\qquad +(-\sin(\alpha)+\cos(\alpha)\partial_1) (\cos(\alpha)+\sin(\alpha)\partial_2)\big)G_{\theta_0,\theta_R}(z,x_0\pm 0,x_0\mp 0), \label{7.22} \\ &M_{\alpha,2,2}(z,x_0,\theta_0,\theta_R) \notag \\ & \quad =\big(-\sin(\alpha) +\cos(\alpha)\partial_1\big)\big(-\sin(\alpha)+ \cos(\alpha)\partial_2\big)G_{\theta_0,\theta_R}(z,x_0,x_0). \label{7.23} \end{align} For relevant references in the context of \eqref{7.9}--\eqref{7.23}, we refer, for instance, to \cite[Ch.\ III]{CL90}, \cite[Ch.\ 9]{CL85}, \cite[App.\ J]{GH03}, \cite{GHSZ95}, \cite{GRT96}, \cite{GS96}, \cite[Ch.\ 2]{Le87}, \cite[Ch.\ 2]{LS75}, \cite[Ch.\ 6]{Pe88}, \cite[Chs.\ II, III]{Ti62}, and the references cited therein. A comparison of equations \eqref{7.1}, \eqref{7.2}, \eqref{7.4}, \eqref{7.5}, \eqref{7.7}, \eqref{7.8} with equations \eqref{7.18}--\eqref{7.20} and \eqref{7.21}--\eqref{7.23}, respectively, clearly exhibits the different character of $\ensuremath{\La_{\te_0,\te_R}} (z)$ and $M_\alpha(z,x_0,\theta_0,\theta_R)$, $\alpha \in [0,\pi)$, despite the fact that both are $2 \times 2$ matrix-valued Herglotz functions whose associated matrix-valued measures contain all spectral information on $\ensuremath{H_{\te_0,\te_R}}$. Additional differences are highlighted in Remark \ref{r3.3}, and we feel that $\ensuremath{\La_{\te_0,\te_R}} (z)$ (and more generally, $\ensuremath{\La_{\te_0,\te_R}^{\te_0^{\prime},\te_R^{\prime}}} (z)$) is deserving of a more detailed study. \medskip We conclude with a final observation: \begin{remark} \label{r6.1} With only minor modifications, all results in this paper extend to general, regular three-coefficient differential expressions of the type \begin{equation} \frac{1}{r} \bigg(- \frac{d}{dx} p \frac{d}{dx} + q\bigg), \quad x\in [0,R], \end{equation} where \begin{equation} p > 0, r > 0 \, \text{ a.e.\ on $(0,R)$}, \quad \frac{1}{p}, q, \, r \in L^1((0,R); dx), \end{equation} generating Sturm--Liouville operator realizations in $L^2((0,R); r dx)$. One just needs to consistently replace the derivative $f'$ for elements in operator domains and Wronskians by the first quasi-derivative $(pf')$. \end{remark} \medskip \noindent {\bf Acknowledgments.} We are indebted to Sergei Avdonin, Pavel Kurasov, and Mark Malamud for helpful discussions on this topic, and to Dorina Mitrea for a critical reading of this manuscript.
\section{Introduction} Stars with spectral type A have long shown evidence for surpising circumstellar disk structures \citep{st84,kgc05,gak06,obh08} and stars with spectral type $\sim$A6 and earlier have become increasingly targeted for low-mass companions through high-contrast imaging \citep{hhs06,hos07,hhk08,oh09} resulting in detections of several low-mass companions \citep{kgc08,mmb08}. Indeed, \citet{jbm07} suggest that the frequency of {\it planet} occurrence around A-type stars is twice that of solar-mass stars. In addition to planetary-mass companions, the frequency and mass ratio distributions of {\it stellar}-mass companions to nearby A stars can help constrain binary formation scenarios---such as models based on the more massive primary star dynamically capturing a lower mass companion \citep{mc93}, or a picture relying on initial fragmentation within a protostellar cloud, e.g. \citet{bb96}. Although some multiplicity studies of A and B stars have been conducted---e.g. the \citet{st02} and \citet{kbz05} surveys of Sco OB2---a comprehensive statistical picture of multiplicity around these massive stars, based on both cluster and field objects, has yet to emerge. Observations of massive, early-type stars may serve as important boundary-type systems, to which models of formation must conform. Specifically, an abundance of brown dwarf/M-dwarf companions to A stars would lend support to recent models describing the formation of these objects through the fragmentation of an initially massive circumstellar disk \citep{kmy09,sw09}. Moreover, the frequency of stellar companions to A-stars may be related to their anomolous source of X-rays. Since A-stars have shallow or non-existent convective regions in their envelopes, they lack a significant dynamo effect, and can be expected to display negligible X-ray emission. Meanwhile, M dwarfs are well known sources of X-rays \citep{fgs93}. \citet{sgh85} suggested that unseen late-type stellar companions to A stars may be the source of the X-rays. Indeed, \citet{ss07} find that the majority of nearby X-ray emitting A-stars have some evidence of possessing low-mass companions, likely responsible for the X-ray emission. Moreover, \citet{zoh10} have recently discovered a mid-M dwarf companion to the nearby A-star Alcor, a ROSAT source. \begin{deluxetable}{llc} \tabletypesize{\scriptsize} \tablecaption{Fundamental Parameters for $\zeta$ Virginis A} \tablewidth{0pt} \tablehead{\colhead{Parameter} & \colhead{} & \colhead{Value}} \startdata Identifiers & & HIP66249, HR5107,\\ & & HD118098 \\ Spectral Type\footnote{\citet{gcg03}} & & A3V \\ $V$ magnitude\footnote{\citet{hj82}} & & 3.40 \\ Parallax (mas)\footnote{\citet{plk97}} & & 44.55$\pm0.90$\\ RA, Dec Proper Motion (mas yr$^{-1}$)\footnote{\citet{plk97}} & & -278.89$\pm0.83$, \\ & & 48.56$\pm0.71$ \\ Radial Velocity (km s$^{-1}$)\footnote{\citet{ab72, ldg09}} & & -13.2$\pm0.3$ \\ \enddata \label{fundpar} \end{deluxetable} \subsection{Previous Studies of $\zeta$ Virginis} The star $\zeta$ (``Zeta'') Virginis (HIP66249, A3V, $V$=3.40---See Table~\ref{fundpar}, hereafter $\zeta$ Vir), is a target in our on-going high-contrast imaging program \citep{odn04,soh07,hob08}. This nearby (22.45 pc) star has been previously used as a calibrator star for interferometric work \citep{acm09}, and has also been followed with the HARPS survey for radial velocity variations \citep{ldg09}. No such variations were found. The Bright Star Catalogue \citep{hj82} lists $\zeta$ Vir as a member of the Hyades moving group. There is some spread in the derived age of this group. Although some authors claim ages as high as 625 Myr \citep{pbl98} or 650 Myr \citep{lfl01}, \citet{rss05} cites the age of this star at 505 Myr. Rather than trying to establish the membership of this star in the Hyades moving group, and then adopt the moving group age for the star, we simply adopt the 505 Myr age of \citet{rss05} for the analysis in this paper. $\zeta$ Vir has indeed been observed by ROSAT \citep{hsv98}, and is listed as a single star, with an X-ray brightness ($L_x=1.07\times 10^{28}$ erg s$^{-1}$). Despite the fact that there is a 20$^{\prime\prime}$ offset between X-ray and optical positions, the ROSAT catalog lists a 14$^{\prime\prime}$ uncertainty on the position of $\zeta$ Vir. Such a 1.5$\sigma$ positional offset easily leaves open the possibility that the observed X-ray source is in fact located at $\zeta$ Vir. The $\zeta$ Vir system has a radial velocity of -13.2 km s$^{-1}$ as mentioned in \citet{ab72} and \citet{kok98} found this radial velocity to be constant at the 1-2 km/s level, i.e. lacking a companion, using it as one of their standard calibrator stars. Speckle observations at the Canada-France-Hawaii Telescope in the optical \citep{mmh93} did not find a binary companion for this star, however this is not surprising given the comparatively low dynamic range ($\Delta M$$\sim$3 mags) of this technique. \citet{pmm01} specifically carried out a survey of bright early-type stars both in the field and in clusters to search for companions using AO, but no mention of this star is listed. \section{Observations} We have imaged the $\zeta$ Vir system using two observing programs, each with different instruments: A coronagraph working together with an infrared camera, and a newly commisioned coronagraph which employs an IFS as the primary science camera. We describe each observing program below. \subsection{``The Lyot Project'', a Coronagraphic Imager at AEOS} The first instrument, ``The Lyot Project'' \citep{odn04,soh07} was a diffraction-limited classical Lyot coronagraph \citep{l39,skm01} working with the Adaptive Optics (hereafter ``AO'') system on the 3.63 m AEOS telescope on Haleakala, Hawaii \citep{rn02}. Our images were gathered using ``Kermit,'' an infrared camera \citep{mdp03}, with a 13.5 mas pixel$^{-1}$ plate scale, and differential polarimetry mode for detection of diffuse circumstellar material \citep{hos09}. Images of $\zeta$ Vir in $J$, $H$, and $K$-bands were obtained using this instrument over three epochs spanning three years as listed in Table~\ref{observations}. Our coronagraph used a focal plane mask with a 455 mas diameter (4.9$\lambda/D$ at $H$-band), as well as its own internal tip/tilt system. Images in the $J$, $H$, and $K$-bands were obtained during the second and third epochs, while only $H$-band was obtained on the first. To calibrate our photometry, we also obtained 1 s unocculted images in addition to the coronagraphically occulted images. In this setup, the target is more than 1$^{\prime\prime}$ away from our occulting mask. The raw data images, both occulted and unocculted, required a mix of both traditional data reduction steps (e.g. dark current subtraction, bad-pixel masking, flat-field correction) as well as some techniques customized for the infrared camera. More details on the data reduction are given in \citet{soh06} and \citet{dho06}. \begin{deluxetable*}{lccccc} \tabletypesize{\scriptsize} \tablecaption{Observations and Astrometric Distance Between Primary and companion for the $\zeta$ Vir System} \tablewidth{0pt} \tablehead{\colhead{Epoch} & \colhead{MJD} & \colhead{Wavelength} & \colhead{Observing Method} & \colhead{Separation (mas)} & \colhead{PA (degrees East of North)} } \startdata 1) & 53168.2839 & $H$-band & Lyot Coronagraph &1846 $\pm$ 9 & 144.7 $\pm$ 0.1 \\ 2) & 53507.3742 & $J$, $H$, $K$-bands & Lyot Coronagraph &1830 $\pm$ 3 & 146 $\pm$ 1.0 \\ 3) & 54257.5378 & $J$, $H$, $K$-bands & Lyot Coronagraph &1814 $\pm$ 16 & 147.4 $\pm$ 0.1 \\ 4) & 54657.1548 & 1.1 - 1.8 $\mu$m (IFS) & APLC + IFS &1790 $\pm$ 12 & 149.8 $\pm$ 0.1 \\ 5) & 54904.4445 & 1.1 - 1.8 $\mu$m (IFS) & APLC + IFS &1779 $\pm$ 12 & 151.0 $\pm$ 0.2 \\ \enddata \label{observations} \end{deluxetable*} \subsection{``Project 1640,'' a coronagraphic Integral Field Spectrograph at Palomar} The second instrument used to image the $\zeta$ Vir system is a new instrument \citep{hob08} recently commissioned on the 200-in Hale Telescope at Palomar Observatory. This instrument, termed ``Project 1640,'' is a coronagraph integrated with an integral field spectrograph (IFS) spanning the $J$ and $H$-bands (1.05$\mu$m - 1.75$\mu$m). The IFS+Coronagraph package is mounted on the Palomar AO system \citep{dbp98}, which in turn is mounted at the Cassegrain focus of the Hale Telescope. The coronagraph is an Apodized-Pupil Lyot coronagraph (APLC) \citep{s05}, an improvement of the classical Lyot coronagraph \citep{skm01}. This coronagraph uses a 370 mas diameter (5.37$\lambda/D$ at $H$-band) focal plane mask. The IFS, or hyper-spectral imager, is a microlens-based spectrograph which can simultaneously obtain $4\times10^4$ spectra across our $4^{\prime\prime}\times4^{\prime\prime}$ field of view. Each microlens subtends 19.2 mas on the sky and a dispersing prism provides a spectral resolution ($\lambda/\Delta\lambda$)$\sim$32. The IFS focal plane consists of $4\times10^4$ spectra that are formed by dispersing the pupil images created by each microlens. To build a data cube, the data pipeline uses a library of images made in the laboratory with a tunable laser, which spans the operating wavelength band of the instrument. Each image contains the response of the IFS to laser emission at a specific wavelength---a matrix of point spread functions. Each laser reference image is effectively a key showing what regions of the $4\times10^4$ spectra landing on the detector correspond to a given central wavelength. Each laser reference image is cross correlated with the focal plane spectra to extract each wavelength channel. The pixel scales for each instrument were calculated at by imaging four binary stars (HIP107354, HIP171, HIP88745, and WDS11182+3132) with high quality orbits with small astrometric residuals \citep{hmw01}. The pixel scale is calculated by performing a least squares fit between these predicted separations and the pixel separation in our data. \section{The Companion} Here we report the discovery of a faint stellar companion to $\zeta$ Vir, hereafter $\zeta$ Vir B. The discovery image is shown in Figure~\ref{zetavir}. To our knowledge, the existence of this companion has not been reported previously. \begin{figure}[ht] \center \resizebox{1.2\hsize}{!}{\includegraphics{fig1.ps}} \caption{A 60 s $H$-band (1.65 $\mu$m) image of the the star $\zeta$ Vir taken on 2004 June 12 (UT) at the AEOS telescope. In this discovery image, the adaptive optics system is on, and the star has been coronagraphically occulted behind our 455 mas occulting mask. A faint stellar companion, $\zeta$ Vir B, 7 magnitudes fainter than the host star and sharing common proper motion with the host star, is apparent at the bottom left of the image.} \label{zetavir} \end{figure} \subsection{Common Proper Motion Analysis} The astrometric measurements for the primary/companion separation are presented in Table~\ref{observations}. For the first and third epochs, the astrometric positions of the two stars were obtained from images in which the primary star was occulted. In these cases, a centroid to each PSF was calculated as part of a best fit radial profile measurement. With coronagraphic imaging, the exact position of the occulted star is difficult to determine. The uncertainty can be estimated using binary stars, in which one of the binary members is occulted \citep{dho06}. For all but the first epoch, we used a physical mask with a superimposed grid \citep{so06}, which produces fiducial reference spots indicating the position of the host star to within $\sim$10 mas. The second, fourth and fifth epochs contained fully unocculted data with sufficiently high signal-to-noise to measure the position of both the primary and the companion. $\zeta$ Vir A, listed as a high-proper motion star, has a proper motion of 283 mas yr$^{-1}$ \citep{plk97}. If these two objects were not associated with each other, we could expect a $\sim$$1.35^{\prime\prime}$ change in separation over the 4.75 year course of observations. Instead, we report a $\sim$200 mas southwesterly change in the position of the companion (see Figure~\ref{zetavirpos}) relative to the host star. Since the relative separation between the host star and the companion is far less than the $\sim$$1.35^{\prime\prime}$ change in separation if the two were not mutually bound, we are confident that these two objects share common proper motion. Moreover, this westerly change reflects the orbital motion of $\zeta$ Vir B over the 4.75 year observing baseline. \begin{figure}[!ht] \center \resizebox{1.0\hsize}{!}{\includegraphics{fig2.ps}} \caption{The offset positions of $\zeta$ Vir B relative to the host star. The position of the host star is marked with the $\star$ symbol. The inset portion of the plot shows the positions of the stellar companion over the 4.75 years of observations presented in this paper. The error bars incorporate the uncertainties in the radial separation and the position angle. The vector labelled ``$\mu$'' at the upper right shows the magnitude and direction of the proper motion of the $\zeta$ Vir system over the 4.75 year duration of these observations (-1325 mas/yr, 230.66 mas/yr). } \label{zetavirpos} \end{figure} \subsection{Photometry} Aperture photometry of the companion was performed on images from the Lyot Project in which $\zeta$ Vir was occulted behind our coronagraphic mask. The flux was summed in apertures of radii of 270 mas, 340 mas, and 300 mas, for the $J$, $H$, and $K$-band images respectively, followed by sky subtraction. Zeropoints for the $J$, $H$ and $K$ data were derived by performing large (760 mas) aperture photometry on unocculted, unsaturated images of $\zeta$ Vir A taken immediately prior to the occulted observations. We calculate $J$, $H$, and $K$-band photometric zero points of $20.990 \pm 0.017$, $20.639 \pm 0.078$, and $19.932 \pm 0.069$, respectively. Assuming a distance of 22.45 pc \citep{plk97}, Table~\ref{phot} shows our calculated absolute $J$, $H$, and $K$-band magnitudes of $8.99\pm .06$, $8.41\pm.14$, and $8.14\pm.17$, respectively for $\zeta$ Vir B. \subsection{Spectroscopy} Integrating an IFS into more conventional high-contrast imaging techniques can provide significantly more information on objects in close vicinity to their host star \citep{sf02,bgf06,mml07}. Normally, when spectra are unavailable, parameters such as mass, spectral type, and age must be derived by combining broadband photomery with model predictions. Such models can be problematic at very young ages \citep{smv06,ajl07}. Observations with our IFS at Palomar Observatory \citep{hob08} allowed us to obtain the spectrum of $\zeta$ Vir B shown in Figure~\ref{spectra}. Each point in the spectrum of $\zeta$ Vir B was calculated by performing aperture photometry on each image in a data cube. Examples of three such images taken from a data cube are shown in Figure~\ref{zetavir1640}. The photometry was obtained using a circular aperture such that the second Airy ring of the Point Spread Function was enclosed at each wavelength in the data cube. A median background sky value was calculated in a 40 mas wide annulus, just outside the photometric aperture, and subtracted from the target counts. Each flux value for $\zeta$ Vir B shown in Figure~\ref{spectra} is a median of five data points taken from five data cubes of $\zeta$ Vir B, and the error bars show the 1$\sigma$ spread of these five values. The spectrum was calibrated using a reference star (HIP 56809, G0V, $V$=6.44) by comparing the measured counts of the reference star with a template G0V star (HD 109358, G0V, $V$=4.26) taken from the IRTF spectral Library \citep{crv05,rcv09} to derive a spectrograph response function. \begin{deluxetable}{ccc} \tabletypesize{\scriptsize} \tablecaption{Photometry for $\zeta$ Virginis B} \tablewidth{0pt} \tablehead{\colhead{Band} & {apparent magnitude} & \colhead{absolute magnitude}} \startdata $J$ & $10.75\pm 0.06$ & $8.99\pm 0.06$ \\ $H$ & $10.17\pm 0.14$ & $8.41\pm 0.14$ \\ $K$ & $9.90\pm 0.17$ & $8.14\pm0.17$ \enddata \label{phot} \end{deluxetable} The spectrum shown in Figure~\ref{spectra} has had this response correction applied to it. We have excluded the data points in the vicinity of the water absorption band between $\sim$1.35 $\mu$m and $\sim$1.5$\mu$m, since the degree of water absorption present in the calibrator star was sufficiently different from that present in the $\zeta$ Vir observations. Also shown in the figure are template spectra for an M2V through M7V star taken from the IRTF spectral Library. The extracted spectrum for $\zeta$ Vir B is most consistent with the M4V - M7V spectral types. \begin{figure}[ht] \center \resizebox{1.15\hsize}{!}{\includegraphics{fig3.ps}} \caption{The $J$ and $H$-band spectrum of $\zeta$ Vir B obtained with our IFS and coronagraph at Palomar \citep{hob08}. Also shown are template spectra for M2V through M7V stars taken from the IRTF spectral Library \citep{crv05,rcv09}. The water band data points between $\sim$1.35 and $\sim$1.5$\mu$m have been excluded due to the variation of this band between observations of $\zeta$ Vir and our calibrator star. } \label{spectra} \end{figure} \begin{figure*} \center \resizebox{1.05\hsize}{!}{\includegraphics{fig4.ps}} \caption{Three images of the $\zeta$ Vir system taken from a data cube from the Project 1640 IFS. The three images show the system at 1.22, 1.55, and 1.70 $\mu$m. The host star, $\zeta$ Vir A has been coronagraphically occulted at the center of each image, and the companion, $\zeta$ Vir B is evident at the lower left. } \label{zetavir1640} \end{figure*} \section{Analysis} \subsection{Mass and Age of $\zeta$ Vir A} Although \citet{hj82} list $\zeta$ Vir as a possible member of the Hyades moving group, several lines of evidence suggest simply assigning the system the age of the Hyades cluster is not rational. Indeed, using criteria based on mass distribution and metallicity, \citet{fpl07} question whether most members of the Hyades Moving Group are actually evaporated members of the Hyades Open Cluster. The mean metallicity of the Hyades cluster has been well established with an [Fe/H] value of $0.14\pm0.05$ \citep{ccl97}, and later refined to $0.144\pm 0.013$ \citep{g00}. However, the mean metallicity value of [Fe/H] $\simeq -0.02$ \citep{gcg03} for the $\zeta$ Vir system is significantly different than the above values, ruling out the possibility that this star is a member of the Hyades cluster. For this work, we have decided to adopt the age of 505 Myr given by \citet{rss05}. In the left panel of Figure~\ref{magmass} we show the luminosity-mass parameter space showing model tracks calculated by \citet{sdf00} for a 505 Myr system with solar metallicity. The vertical extent of the box indicates the absolute $V$-band photometric uncertainty. This uncertainty in the luminosity ($M_V = 1.64\pm 0.05$), defines an allowable mass region for $\zeta$ Vir A of $2.041\pm .024$ M$_\odot$. \begin{figure*}[ht] \center \resizebox{1.0\hsize}{!}{\includegraphics{fig5.ps}} \caption{The left hand panel shows the luminosity-mass relation for the host star, $\zeta$ Vir A as calculated by \citet{sdf00} for a 505 Myr system. We use absolute $V$-magnitude in lieu of total luminosity. The vertical extent of the box indicates the $V$-band photometric uncertainty ($M_V = 1.64\pm 0.05$) for $\zeta$ Vir A. This defines an allowable region for the mass for the A3V host star to be $2.041\pm .024$ M$_\odot$. The right panel shows evolutionary models for low-mass stars taken from \citet{bcb03} for the $J$, $H$, and $K$-bands. As with the case for $\zeta$ Vir A, the vertical extent of each box indicates the photometric uncertainty at each band. These three bandpass values gives an overall value of $0.168^{+.012}_{-.016}$ M$_\odot$ for $\zeta$ Vir B. } \label{magmass} \end{figure*} \subsection{Mass and Age of $\zeta$ Vir B} We use the models of \citet{bcb03} to derive a mass for $\zeta$ Vir B assuming the age of 505 Myr. In Figure ~\ref{magmass} (right panel) we show plots of the $J$, $H$, and $K$-band absolute magnitudes for a range of companion masses calculated from these models. As with the case of $\zeta$ Vir A, the uncertainty in the photometry of this object has very little effect on the derived mass of the companion. Together these values give an overall derived mass of $0.168^{+.012}_{-.016}$ M$_\odot$. We take this value as the final derived mass for $\zeta$ Vir B. To check the validity of this model-based mass estimate, we compare this value with empirically derived mass-luminosity relations given in \citet{hm93} and \citet{dfs00}. Using the $J$, $H$, and $K$ magnitudes, these two works predict values of $0.152 \pm 0.009$ M$_\odot$ and $0.166 \pm 0.004$ M$_\odot$ (See Table~\ref{masstable}), consistent with a mid-M spectral type for a main sequence star at these ages \citep{rhg95,hgr96}, and consistent with the spectral determination derived previously. The \citet{hm93} and \citet{dfs00} mass values are slightly lower than the model-based 0.168 M$_\odot$ value given by the \citet{bcb03} models, but are still consistent with a mid-M spectral type for $\zeta$ Vir B. Using the range of primary star masses derived above gives this system a mass ratio of $q = 0.082^{+.007}_{-.008}$. \begin{deluxetable*}{llcccc} \tabletypesize{\scriptsize} \tablecaption{$J$,$H$,and $K$-band Mass Determinations for $\zeta$ Vir B.} \tablewidth{0pt} \tablehead{\colhead{Model} & \colhead{Method} & \colhead{$J$-band} & \colhead{$H$-band} & \colhead{$K$-band} & \colhead{Median$\pm$1$\sigma$}} \startdata \citet{hm93} & Mass-Luminosity (Empirical) & 0.136 M$_\odot$ & 0.152 M$_\odot$ & 0.152 M$_\odot$ & $0.152\pm 0.009$ M$_\odot$ \\ \citet{bcb03} (505 Myr) & Evolutionary (Theoretical) & 0.168 M$_\odot$ & 0.169 M$_\odot$ & 0.167 M$_\odot$ & $0.168\pm 0.001$ M$_\odot$ \\ \citet{dfs00} & Mass-Luminosity (Empirical) & 0.170 M$_\odot$ & 0.166 M$_\odot$ & 0.163 M$_\odot$ & $0.166\pm 0.004$ M$_\odot$ \\ \enddata \label{masstable} \end{deluxetable*} \subsection{Orbital Analysis} \begin{figure*}[ht] \center \resizebox{.9\hsize}{!}{\includegraphics{fig6.ps}} \caption{The loci of possible values of the semi-major axis (left), and eccentricity (right), for $\zeta$ Vir B, assuming a range of line-of-sight positions and velocities \citep{gbk98}. The contours for the left plot are 27, 35, 60, 100, 160, and 300 AU. The right hand plot shows eccentricity contours equal to 0.25 0.50, 0.60, 0.75, and 0.90. This constrains the semi-major axis to be $\gtrsim 24.9$ AU (and hence the period to be $\gtrsim124$ yrs), and eccentricity $\gtrsim 0.16$.} \label{orbconstrain} \end{figure*} Given the relatively short span (4.75 years) of the observations of $\zeta$ Vir B (see Figure~\ref{zetavirpos}), fitting an orbital model to the data is premature. However, we may borrow an analysis used by \citet{gbk98} to constrain the eccentricity, $e$, and semi-major axis, $a$, of $\zeta$ Vir B using our astrometry in the two-dimensional plane of the sky, combined with Kepler's Laws. We refer the reader to \citet{gbk98} for a full explanation. Over our 4.75 year time baseline, we calculate a velocity of -1.139 AU yr$^{-1}$ in the westward direction, and .062 AU yr$^{-1}$ in the south direction. Assuming the orbit of $\zeta$ Vir B is bounded, and assuming a range of line-of-sight positions and velocities for $\zeta$ Vir B, we are able to constrain $a$ and $e$, and we show the loci of possible values for these parameters in Figure~\ref{orbconstrain}. The ordinate and abcissa values show the assumed values of the line-of-sight velocity, and position, respectively. The values shown in Figure~\ref{orbconstrain} indicate $a\gtrsim 24.9$ AU and $e\gtrsim 0.16$. From the semi-major axis constraint, we can constrain the period $P=(4\pi^2a^3/\mu)^{1/2}$ to be $\gtrsim 124 $ yrs, where $\mu = G(m_1 + m_2)$. In this analysis we have assumed a mass of $m_1=2.04 M_\odot$ (see Figure~\ref{magmass}) and $m_2 = .168 M_\odot$ for $\zeta$ Vir A and $\zeta$ Vir B, respectively. \section{Discussion} Assuming a mass of $2.041\pm .024$ M$_\odot$ for $\zeta$ Vir A gives this newly discovered binary system a mass ratio of $q = 0.082^{+.007}_{-.008}$. Although numerous low-mass companions have been detected around $\sim$1 M$_\odot$ stars, this system is of particular interest given that it orbits a primary star of $\sim$2 M$_\odot$. The $q$ distributions for primaries in the mid and low mass stellar regimes have been well studied, but comprehensive binary statistics for A stars are incompletely surveyed. The mass ratio distribution for the mid and lower mass ranges show fairly clear trends with mass. Namely, \citet{brs07} discusses that very low mass binaries have mass ratios that are skewed towards unity. Studies using a complete sample of stars between 0.6 - 0.85 M$_\odot$ \citep{msp03}, as well as M-dwarf surveys \citep{fm92,rg97}, find a significantly flat distribution of mass ratios. In the same vein, \citep{kim08}, used a sample of 82 young stars of GKM type in the Upper Sco star forming region and found a distribution of mass ratios not significantly different from a constant distribution, i.e. not significantly biased towards having low mass companions. Towards higher masses, \citet{dm91} find that F and G type stars have a broad range of mass ratios, but with a slight increase towards small secondary masses. Finally, at the higher mass end, \citet{st02} and \citet{kbz05} have performed a survey of A and B-stars in the Sco OB2 association, finding a high rate of binarity. Similarly, many studies show a positive correlation between the distribution of binary separations and the mass of the system. The \citet{dm91} sample of solar mass stars show a mean separation of $\sim$30 AU. At lower masses, the \citet{fm92} survey (largely M dwarfs within 8 pc) and \citet{rg97} surveys (M dwarfs within 20 pc) find mean separations between 4 and 30 AU. Continuing towards lower mases, low mass M dwarfs and brown dwarfs as discussed in \citet{brs07}, have notably smaller mean separations ($\sim$4 AU), with maximum separations $\sim$20 AU. \citet{kbz05} and \citet{st02} have undertaken the first steps towards a comprehensive study of the multiplicity of A stars. Our work aims to aid in that effort. The value of the current study lies in the ability to obtain a spectrum which determines the spectral type, a tight constraint on the secondary mass, and a constrained orbit. Our finding may also be useful for studies of the anomalous X-ray emission from A stars. Unseen late-type companions to A stars have been hypothesized to be the source of their anomolous X-ray emission. Indeed, the presence of the M-dwarf companion in this system can easily account for the X-ray flux detected by ROSAT. For a $.168 M_\odot$ star at 505 Myr, the \citet{bcb03} evolutionary tracks predict a luminosity of $L_{bol} \simeq 2\times10^{31}$ erg s$^{-1}$. And assuming that the X-ray luminosity $L_{x} = 1.07\times10^{28}$ erg s$^{-1}$ noted by \citet{hsv98} is due completely to the companion, this predicts a $\log(L_{x}/L_{bol})$ value of -3.3, quite typical for a young mid M dwarf (See \citet{fgs93}, especially their Figure 3). Recently, \citet{zoh10} have reported the presence of a mid-M dwarf bound to the star Alcor. More complete high-contrast surveys for companions surrounding an ensemble of A stars will allow researchers to begin to address the issue of the anomolous X-ray emission in a statistically robust manner. An abundance of M-dwarf companions in configurations like this also may lend support to models of binary formation based on fragmentation of a massive circumstellar disk. As \citet{kmy09} point out, massive stars with their presumably massive circumstellar disks and correspondingly high mass infall rates provide an environment conducive for the formation of disk instabilities and fragmentation. Indeed, such a mechanism is more likely for disks surrounding stars more massive than 1-2M$_\odot$\citep{kkm07,kmk08}. If indeed this fragmentation is a prominent mechanism for the formation of binary companions \citep{sw09}, the abundance of low mass companions (brown dwarfs and M-dwarfs) should be more frequent around more massive stars. \section{Summary} We report the discovery of a low-mass, M4V-M7V stellar companion to the star $\zeta$ Vir. This object clearly shares common proper motion with its host star, and we derive a mass of $0.168^{+.012}_{-.016}$ M$_{\odot}$, corresponding to a mass ratio $q = 0.082^{+.007}_{-.008}$. Our broad-band photometry and spectroscopy are consistent with an mid-M spectral type. Although numerous low-mass companions have been identified around $\sim$1 M$_\odot$ systems, this object is significant given its membership in a $\sim$2 M$_\odot$ system. Characterization of more systems like this are important for identifying the anomalous source of X-rays from A stars as well as constraining possible modes of formation of stellar companions through the fragmentation of massive circumstellar disks. \acknowledgments We thank the anonymous referee for his or her comments. This work was performed in part under contract with the California Institute of Technology (Caltech) funded by NASA through the Sagan Fellowship Program. The Lyot Project is based upon work supported by the National Science Foundation under Grant Nos. AST-0804417, 0334916, 0215793, and 0520822, as well as grant NNG05GJ86G from the National Aeronautics and Space Administration under the Terrestrial Planet Finder Foundation Science Program. A portion of the research in this paper was carried out at the Jet Propulsion Laboratory, California Institute of Technology, under a contract with the National Aeronautics and Space Administration and was funded by internal Research and Technology Development funds. The Lyot Project grateful acknowledges the support of the US Air Force and NSF in creating the special Advanced Technologies and Instrumentation opportunity that provides access to the AEOS telescope. Eighty percent of the funds for that program are provided by the US Air Force. This work is based on observations made at the Maui Space Surveillance System, operated by Detachment 15 of the U.S. Air Force Research Laboratory Directed Energy Directorate. This work has been partially supported by the NSF Science and Technology Center for Adaptive Optics, managed by the University of California at Santa Cruz under cooperative agreement AST 98-76783. The Lyot Project is also grateful to the Cordelia Corporation, Hilary and Ethel Lipsitz, the Vincent Astor Fund, Judy Vale and an anonymous donor, who initiated the project.
\section*{Prologue} It has been 50 years since Joseph Weber first embarked on a serious experimental program to try to detect gravitational waves directly~\cite{weber60}, motivated by the possibility of detecting signals from sources such as a core-collapse supernova or a binary neutron star~\cite{weber66}. The intervening years have seen great advances in technologies and new techniques for detecting gravitational waves, from much-improved ``Weber bars'' to highly sensitive broadband interferometers, Doppler tracking of spacecraft such as Cassini~\cite{CassiniGW}, long-term campaigns to monitor pulse arrival times from stable pulsars~\cite{PulsarTiming}, and mature plans for long-baseline interferometer networks in space (namely LISA~\cite{LISA} and DECIGO~\cite{DECIGO}). In parallel, after the discovery of the first binary pulsar in 1974~\cite{HulseTaylor}, radio pulse timing campaigns with a number of short-period binary pulsars have provided compelling ``indirect'' evidence for the existence of gravitational radiation as well as precise experimental tests of the general theory of relativity~\cite{DoublePulsar}. Theoretical work and numerical modeling have provided a much better understanding of the likely gravitational-wave (GW) signatures of the original leading source candidates---core-collapse supernovae~\cite{OttSNGW} and binary neutron stars~\cite{Blanchet35}---as well as many other expected or plausible GW sources, including binary systems with supermassive or stellar-mass black holes, short-period white dwarf binaries, non-axisymmetric spinning or perturbed neutron stars, cosmic strings, early-universe processes, and more. (General overviews of GW sources may be found in \cite{CutlerThorne} and \cite{SchutzOverview}, for instance.) And here is what our direct searches have yielded so far: {\em Nothing}. The lack of a directly detected signal is not surprising, based on our limited knowledge of source populations and on the current sensitivity levels of the detectors. Further instrumental improvements are on the way, including substantial upgrades to the current large interferometers, proposals to build additional interferometers, pulse timing measurements of more pulsars for longer time spans with better precision, and eventually the launch of space missions to open up the low-frequency window that is certain to be rich with signals. According to current schedules, we are sure to be detecting signals and doing GW astronomy around the middle of the coming decade. However, in this article I will argue that we are {\em already} doing GW astronomy. In the next section I will summarize and interpret several completed searches for which the lack of a detectable signal provides some relevant information about the population and/or astrophysics of plausible sources. I will then project forward to the time when signals {\em are} detected and discuss what we may learn from them, and how they will fundamentally change the field of GW astronomy. \section{The impact of null results published so far} In 1969 Weber claimed that his detectors had produced definitive evidence for the discovery of gravitational waves~\cite{weber69}, the first in a series of claims by him that could not be reproduced by others and were ultimately discredited~\cite{collins}. Nevertheless, his attempts inspired an experimental community that continued to improve the detection technologies, cooling bars to cryogenic temperatures to minimize thermal noise and exploring other detection methods~\cite{Thorne300yrs}. Large cryogenic bar detectors---ALLEGRO~\cite{ALLEGRO}, EXPLORER~\cite{EXPLORER}, NAUTILUS~\cite{NAUTILUS}, NIOBE~\cite{NIOBE}, and AURIGA~\cite{AURIGA}--- began operating for extended periods with good sensitivity in the 1990s. Analyzing the data from two or more detectors together substantially reduced the false alarm rate from spurious signals in the individual detectors (from mechanical vibrations, cosmic rays, etc.)\ and enabled searches for weaker and less-frequent transient signals. This culminated in the 1997 formation of the International Gravitational Event Collaboration (IGEC)~\cite{IGECoper}. Around the same time, GW searches with prototype interferometric detectors ({\it e.g.}~\cite{BNS40m}) gave way to the commissioning and eventual operation of the full-scale interferometers TAMA\,300~\cite{TAMA}, LIGO~\cite{LIGO}, GEO\,600~\cite{GEO} and Virgo~\cite{Virgo}, which have by now surpassed the bars in searching for ``high-frequency'' GW signals (above $\sim10$~Hz). Also, the past several years have seen advances in using multiple millisecond pulsars to search for GW signals at frequencies around $10^{-9}$--$10^{-7}$~Hz, for instance with the Parkes Pulsar Timing Array project~\cite{PPTA}. Over the past decade, dozens of papers have been published to report results from searches for various types of GW signals. Aside from a few hints of excess event candidates that were not confirmed, all searches so far have yielded null results. From these are derived upper limits on the rate and/or strength of GW signals reaching the detectors, or alternatively on the possible population of sources. In this section I highlight several of these results which represent significant steps toward gravitational-wave astronomy. Many of these signals are expected to be detectable by the current instruments only if they originate relatively nearby; therefore we begin by exploring our cosmic neighborhood. \subsection{Getting to know our neighbors better} Our galaxy is thought to contain $10^8$--$10^9$ neutron stars~\cite{NarayanOstriker}, of which a few thousand have been detected as radio or X-ray pulsars. A rapidly spinning neutron star can emit periodic gravitational waves through a number of mechanisms, including a static deformation that breaks axisymmetry~\cite{JKS}, persistent matter oscillations ({\it e.g.\ r}-modes)~\cite{Andersson}, or free precession~\cite{VDBprecession}. The Crab pulsar, at a distance of about 2 kpc, is a particularly interesting neighbor. With a current spin frequency of $29.7$~Hz and spin-down rate of $-3.7 \times 10^{-10}$~Hz/s~\cite{CrabMonthly}, its energy loss rate is estimated to be $4 \times 10^{31}$~W. This powers the expansion and electromagnetic luminosity of the Crab Nebula, but the energy flows in this complex system are difficult to pin down quantitatively~\cite{CrabRev2008}, leaving open the possibility that a significant fraction of the energy could be emitted as gravitational radiation. Palomba has estimated that the observed braking index of the pulsar spin-down constrains this fraction to be no more than 40\%~\cite{Palomba}. The LIGO Scientific Collaboration (LSC) used data from the first 9 months of LIGO's S5 science run to search for a GW signal from the Crab pulsar and, finding none, were able to set an upper limit of 8\% of the total spin-down energy, using X-ray image information to infer the orientation of the pulsar spin axis and assuming that the GW emission is phase-locked at twice the radio pulse frequency~\cite{Crab9month}. Analysis of the full S5 data~\cite{KnownPulsarsS5} has improved this limit to just 2\%. This observational result directly constrains the properties of the Crab pulsar and the energy balance of the nebula. The LSC and the Virgo Collaboration (now analyzing data jointly) have also searched for periodic GWs from all known pulsars with spin frequencies greater than 20~Hz and sufficiently precise radio or X-ray pulse timing to allow a fully-coherent search, again assuming that the GW emission is phase-locked at twice the pulse frequency. The analysis~\cite{KnownPulsarsS5} considered 115 radio pulsars (including 71 in binary systems) that were timed during the LIGO S5 run by the Jodrell Bank Observatory, the Green Bank Telescope, and/or the Parkes radio telescope, along with the X-ray pulsar J0537$-$6910 which was monitored by the Rossi X-ray Timing Explorer (RXTE). For each pulsar, a 90\% upper limit was placed on the GW amplitude in terms of the parameter $h_0$, which represents the strain amplitude that would reach the Earth in the ``plus'' and ``cross'' polarizations {\em if} the pulsar spin axis were oriented optimally. Figure~\ref{fig:CWResultsPlot} shows these upper limits, which are remarkably small numbers in themselves. \begin{figure}[bt] \begin{center} \includegraphics[width=11cm]{figure1.eps} \caption{Upper limits on GW strain amplitude parameter $h_0$ for known pulsars using data from the LIGO/GEO\,600 S3, S4 and S5 science runs, taken from~\cite{KnownPulsarsS5} (reproduced by permission of the AAS). Each symbol represents one pulsar and is plotted at the expected GW signal frequency (twice the spin frequency). The grey band is the range expected for S5 given the average instrumental noise level.} \label{fig:CWResultsPlot} \end{center} \end{figure} The lowest upper limit is $2.3 \times 10^{-26}$, obtained for pulsar J1603$-$7202. J0537$-$6910 is the only pulsar besides the Crab for which the upper limit from this analysis reaches the spin-down limit, assuming that the moment of inertia is within the favored range of (1 to 3) $\times 10^{38}$~kg\,m$^2$. Assuming the neutron stars to be triaxial ellipsoids, these amplitude limits may be re-cast as limits on the equatorial ellipticity $\varepsilon$. Pulsar J2124$-$3358, at a distance of $\sim$200~pc and GW frequency of 404~Hz, yields the strictest upper limit, $\varepsilon < 7.0 \times 10^{-8}$. One may ask whether the perfect or near-perfect axisymmetry of these neutron stars is due to the properties of the neutron star material. It has long been thought that conventional neutron stars could support an ellipticity up to a few times $10^{-7}$~\cite{UCB}, but recent work suggests that the pressure in the crystalline crust suppresses defects~\cite{HorowitzKadau} and ellipticities of up to $\sim 4 \times 10^{-6}$ are possible in a conventional neutron star. Most of the pulsars in the S5 analysis have $\varepsilon$ limits below that level, meaning that these neutron stars, at least, are closer to axisymmetric than required by the intrinsic material properties. \subsection{When the neighbors are disturbed...} Soft gamma repeaters (SGRs) are believed to be magnetars, {\it i.e.}\ neutron stars with very strong magnetic fields~\cite{DuncanThompson}. SGRs are observed to emit intense flares of soft gamma rays at irregular intervals which may be associated with ``starquakes'', abrupt cracking and rearrangement of the crust and magnetic field. These events could excite quasinormal modes of the neutron star which then radiate gravitational waves~\cite{dFPSGRGW}. The fundamental mode, at a frequency of around $1.5$ to 3~kHz, is expected to be the most efficient GW emitter, although other nonradial modes may also participate. The LSC have used LIGO data to search for GW bursts associated with flares of SGRs 1806$-$20 and 1900$+$14, including the December 2004 giant flare of SGR 1806$-$20~\cite{Horvath}. A first analysis~\cite{S5y1SGR} treated the flares individually, setting upper limits on GW energy as low as a few times $10^{45}$~erg, depending strongly on the waveform assumed. The best of these limits (for signals in the most sensitive range of the instruments, 100--200~Hz) are within the range of possible GW energy emission during a giant flare, $10^{45}$ to $10^{49}$~erg, according to modeling by Ioka~\cite{IokaSGR}. Unfortunately, the giant flare occurred between LIGO science runs, and the less-sensitive data available at that time only yields upper limits on GW energy emission of $5 \times 10^{47}$~erg and above. A later paper re-analyzed the ``storm'' of SGR 1900+14 flares that spanned a period of $\sim$30 seconds on 29 March 2006~\cite{SGRstack}. This analysis ``stacked'' the data around the times of the individual flares in order to gain sensitivity under the assumption that many or all of the flares had an associated GW burst at a common relative time offset. Two scenarios were considered to choose the relative weighting of the flares: one in which the GW burst energy is assumed to be proportional to the electromagnetic fluence of each flare, and the other in which all large flares are assumed to have more-or-less equal GW burst energy. This analysis yielded per-burst energy upper limits as low as $2 \times 10^{45}$~erg, an order of magnitude lower than the limits set for this storm by the earlier single-burst analysis. These searches are just beginning to address the few existing models of GW emission by SGRs, and are motivating new modeling of SGRs and their disturbances. Stronger constraints (if not a detection) will be obtained when another giant flare occurs while the GW detector network is operating with good sensitivity, and/or from searches using ordinary flares from closer SGRs such as the recently discovered SGRs J0501$+$4516~\cite{Rea0501,Aptekar0501} and J0418$+$5729~\cite{vdH0418}, which may both be less than 2~kpc away. \subsection{Listening for invisible neighbors} As noted previously, only a small fraction of the hundreds of millions of neutron stars in our Galaxy are visible to us in radio waves, X-rays or gamma rays. It is quite plausible that one or more nearby, unseen neutron stars have a large enough asymmetry and spin rate to be emitting periodic gravitational waves at a detectable level. A general argument, originated by Blandford and extended in a 2007 paper by the LIGO Scientific Collaboration~\cite{S2Fstat}, starts with the (very optimistic) assumptions that all neutron stars are born with a high spin rate and spin down due to GW emission alone, and concludes that the strongest signal that we can expect (in an average sense) to receive on Earth would have $h_0 \simeq 4 \times 10^{-24}$, independent of frequency and $\varepsilon$. Knispel and Allen have greatly refined this analysis~\cite{KnispelAllen}, replacing Blandford's simple assumptions about the neutron star population with a detailed simulation of the birth, initial kick and subsequent motion of neutron stars in the Galaxy. For a nominal ellipticity of $10^{-6}$, they find that the maximum expected GW signal amplitude (with the same optimistic assumptions about neutron stars spinning down due to GW emission) is around $10^{-24}$ over the frequency range 100--1000~Hz. The LSC have published two all-sky searches for periodic GW signals using data from the early part of the S5 run, one using the semi-coherent ``PowerFlux'' method~\cite{EarlyS5PowerFlux} and the other using the substantial computing power provided by the Einstein@Home project to carry out longer coherent integrations on a smaller data set~\cite{EarlyS5EatH}. These searches had comparable sensitivities, both slightly surpassing the Knispel and Allen model expectations (with $\varepsilon=10^{-6}$) for pulsars with favorable orientations and GW signal frequencies in the vicinity of 200~Hz. Thus, periodic GW searches may be on the verge of detecting unseen neutron stars, or at least constraining models for the population of such objects in our Galaxy. \subsection{Listening to the galaxy next door} On 1 February 2007, an extremely intense gamma-ray burst was detected by detectors on the Konus-Wind, INTEGRAL, MESSENGER, and Swift satellites. The initial position error box from the relative arrival times of the bursts~\cite{GRB070201loc} intersected the spiral arms of M31 (the Andromeda galaxy), raising the intriguing possibility that it originated in that galaxy, only $\sim$770~kpc away. Furthermore, the leading model for most short-hard GRBs such as this one is a binary merger involving at least one neutron star~\cite{NakarSHGRB}. Such an event would also emit strong gravitational waves. At the time of the GRB, the two detectors at the LIGO Hanford Observatory were collecting science-mode data, while the other large interferometers were not. The LSC searched this data for both an inspiral signal leading up to the merger and for an arbitrary GW burst associated with the merger itself~\cite{GRB070201search}. No plausible signal was found, and from the absence of a detectable inspiral signal at that range, a compact binary merger in M31 was ruled out with $>99\%$ confidence. The LIGO null result, along with a refined position estimate for the GRB~\cite{Mazets070201}, helped to solidify the case that this was most likely an SGR giant flare event in M31~\cite{Ofek070201}. \subsection{Checking out a monster sighting} In 2003 a team of radio astronomers reported evidence for the discovery of a supermassive black hole binary in the bright radio galaxy 3C~66B~\cite{Sudou}, which is located about 90~Mpc from Earth. Their claim was based on very long baseline interferometry (VLBI) observations of the galaxy, in which the radio core of the galaxy was seen to move slightly over the course of 15 months in a manner consistent with an elliptical orbit with a period of 1.05 year. This suggested the presence of a binary system with a total mass near $5 \times 10^{10}\,M_\odot$. Remarkably, such a system would also be expected to merge in $\sim$ 5 years due to energy loss by GW emission. Jenet, Lommen, Larson \& Wen determined that pulsar timing could be used to check this claim, since the gravitational waves from the binary would cause pulse time-of-arrival variations of up to several microseconds~\cite{Jenet3C66B}. They analyzed seven years of archival Arecibo timing data for PSR 1855+09 and found no such variation, definitively ruling out the proposed binary system in 3C~66B. \subsection{Catching a merger anywhere in the sky} Binary neutron star systems are benchmark sources for ground-based GW detectors because binary pulsars give a glimpse of the population (see discussion in \cite{KalogeraBNS}) and efficient GW emission during the inspiral phase just before merging makes them detectable out to considerable distances, currently tens of megaparsecs. Black-hole-and-neutron-star (BHNS) and binary black hole (BBH) systems with stellar-mass black holes can be detected at even greater distances; population synthesis studies suggest that the net detection rates for those sources are likely to be comparable~\cite{OSNSBS,KalogeraBBH}. Because GW detectors have wide antenna patterns, signals from these events can be detected from essentially anywhere in the sky and at any time. The most sensitive search published to date for these sources used LIGO S5 data and templates for binary inspirals with total mass up to $35\,M_\odot$~\cite{S5CBClow}. No significant signal candidate was detected. The LSC interpreted this null result using a population model based on the assumption that the rate of mergers in each nearby galaxy is proportional to its blue light luminosity, as a tracer of massive star formation. The upper limits obtained from a total of 18 calendar months of LIGO data, in units of merger rate per year per $L_{10}$ (defined as $10^{10}$ times the blue light luminosity of the Sun), were $0.014$, $0.0036$ and $0.00073$ for binary neutron star, BHNS, and BBH systems, respectively, calculated assuming black hole masses of $5 \pm 1\,M_\odot$. These limits are still far from the theoretically expected rates, but are motivating the numerical relativity community to improve waveform calculations and the data analysis community to find better ways to search for inspirals with higher masses, significant spin and/or high mass ratio~\cite{EOBNR,EOBresum}. \subsection{Being vigilant for arbitrary bursts} Supernova core collapse and several other plausible signals are not modeled well enough to use matched filtering, either because the astrophysics is not completely known or because the physical parameter space is too large to effectively cover with a template bank. Even for the important case of BBH mergers, which numerical relativity calculations are now having considerable success in modeling~\cite{NRadvances}, the physical parameter space allows for a wide variety of waveforms. Thus it is important to search for arbitrary transient signals (bursts) in the GW data, and data analysis methods have been implemented which robustly detect a wide range of signals without advance knowledge of the waveform. So far the IGEC network of bar detectors is the observation time champion, having collected enough data since 1997 to establish an upper limit of $1.5$ per year on the rate of strong GW bursts~\cite{IGECburst}, with looser rate limits on weaker bursts. More recent IGEC data extended their sensitivity to somewhat weaker bursts~\cite{IGEC2burst}, but with looser rate limits of $\sim$8.5 per year. On the other hand, LIGO is the sensitivity champion. The latest published burst search results~\cite{S5y1Burst}, from the first calendar year of the LIGO S5 run, set upper limits on the rate of bursts arriving at Earth as a function of signal waveform and amplitude, expressed as the root-sum-squared GW strain calculated from both polarization component amplitudes at the Earth: \begin{equation} h_{\textrm{rss}} = \sqrt{{\int_{-\infty}^{+\infty} \left( |h_+(t)|^2 + |h_{\times}(t)|^2 \right) \, dt}} \,\,. \end{equation} Figure~\ref{fig:BurstPlotSG} \begin{figure}[bt] \begin{center} \includegraphics[width=11cm]{figure2.eps} \caption{Burst rate versus strength limit plot for sine-Gaussian waveforms~\cite{S5y1Burst}. The area above each curve is excluded at 90\% confidence level. Results are shown from the LIGO S1, S2 and S4 science runs in addition to the first calendar year of the S5 run (labeled ``S5'' here), showing the progression of increasing sensitivity (to the left) and observation time (downward). Reprinted figure 8a with permission from B.\ P.\ Abbott {\it et al}, Physical Review D {\bf 80}, 102001 (2009). Copyright (2009) by the American Physical Society.} \label{fig:BurstPlotSG} \end{center} \end{figure} shows the limits set by burst searches in the S5 run (first calendar year only) and earlier science runs for ``sine-Gaussian'' waveforms with $Q=9$ and central frequencies up to $\sim$1~kHz. The area above each curve is excluded at 90\% confidence level, {\it i.e.}\ the curve traces out the 90\% upper limit on the rate for a given waveform assuming a hypothetical population with fixed $h_{\textrm{rss}}$. Sufficiently loud bursts of any form have the same rate limit, $3.75$ per year, as the efficiency of the analysis pipeline approaches unity. The signal strength can also be related in a robust way to GW energy emission from a source at a known or assumed distance $r$. For instance, the S5 first-year search mentioned above would have been sensitive to an event in the Virgo galaxy cluster ($r \simeq 16$~Mpc) that emitted $\sim 0.25\,M_\odot c^2$ of GW energy in a burst with a dominant frequency of $\sim 150$~Hz \cite{S5y1Burst}. These searches constrain populations of sources such as binary mergers of intermediate-mass black holes, although so far only preliminary quantitative studies have been made with realistic simulated waveforms~\cite{ninja}. \subsection{Tuning in to spacetime shivering} The Big Bang may have left behind a stochastic background of gravitational waves, isotropic like the cosmic microwave background (CMB) but carrying information about much earlier fundamental processes in the early universe; see \cite{Maggiore} for a review and references in \cite{S5stoch} for updates on the details of plausible processes. A stochastic background, isotropic or not, can also be produced by a large number of overlapping astrophysics sources such as binary mergers, cosmic (super)strings, or core-collapse supernovae. The GW signal generally has the form of random ``noise'' with a characteristic power spectrum, though it can be distinguished from true instrumental noise by testing for a common signal in multiple detectors for which the instrumental noise is known to be uncorrelated. Jenet {\it et al} have used pulsar timing to search for low-frequency stochastic gravitational waves in archival and newly-obtained data for seven pulsars spanning intervals from $2.2$ to $20$ years~\cite{PTAstoch}. They detected no signal but placed limits on the GW energy density assuming different power-law distributions as a function of frequency. From these they also derive limits on mergers of supermassive binary black hole systems at high redshift, relic gravitational waves amplified during the inflationary era~\cite{Grishchuk2005}, and a possible population of cosmic (super)strings~\cite{DamourVilenkin}. The LSC and Virgo have used LIGO data to search for a stochastic GW signal in the vicinity of 100~Hz by measuring correlations in the data from the Hanford and Livingston interferometers to test for a signal well below the noise level of either instrument. A recently published paper~\cite{S5stoch} used the data from the full S5 science run to set a limit on the energy density in GWs as a fraction of the critical energy density needed to close the universe. The result, assuming a frequency-independent spectrum, was $\Omega_0 < 6.9 \times 10^{-6}$ at 95\% confidence. This direct limit surpasses the indirect limits from Big Bang nucleosynthesis and the CMB and constrains early-universe models. It also imposes constraints on a possible population of cosmic strings in a different part of the parameter space than the pulsar timing result does. \subsection{Summary: impact of null results} From this sampling of search results, most published in the past few years, one can see the beginnings of rich astrophysics coming out of GW observations. The searches are now placing meaningful constraints on some individual objects and events, source populations (either real or speculated), and the total energy density of gravitational waves in the universe. Many more analyses are in progress, and null results will surely continue to provide interesting information. \section{The (future) impact of detected signals} As I write this sentence, I can see\footnote{I checked the web page http://www.ldas-sw.ligo.caltech.edu/ligotools/runtools/gwistat/, which reports the current status of operating GW detectors, on 29 November 2009.} that the two LIGO 4-km detectors and Virgo are all collecting science-mode data at this particular moment (as part of the ongoing S6/VSR2 science run), while AURIGA, EXPLORER and NAUTILUS are also collecting good data. GEO\,600 is being upgraded to ``GEO-HF'' with a focus on improving the sensitivity for frequencies above $\sim$400~Hz~\cite{GEO-HF} and will collect more data over the next several years. It is possible that the first unambiguous GW signal is in the data already collected but not yet fully analyzed, or will soon be recorded. Besides proving without a doubt that GWs exist and can be detected, even a single detection would give us invaluable information about the source from the waveform properties. In the case of a binary inspiral, the ``chirp'' rate and possible modulation reflect the component object masses and spins; for the ringdown of a perturbed black hole, the damped-sinusoid frequency and decay rate reveal the mass and spin; for a spinning neutron star, the signal amplitude and polarization content indicate its ellipticity and spin axis inclination; and so on. A signal that does not match any of the standard models could confirm a speculative source type or reveal an unanticipated one. The reconstructed sky position of the source may point to a galaxy and thus pin down the distance. If the signal is associated with an astronomical event or object observed by other means---such as a GRB, optical or radio afterglow, supernova, neutrino detection, or known pulsar---then the complementary information will provide a clearer view of the nature of the source and emission mechanisms. Preparations are well underway for major upgrades to the ground-based GW detector network in the form of Advanced LIGO~\cite{AdLIGO} and Advanced Virgo~\cite{AdVirgo}, which will have an order of magnitude better sensitivity than the current instruments. When these detectors begin operating in 2014 or 2015, we can expect regular detections of binary inspirals and excellent prospects for detecting various other signals. The detection of multiple signals of the same type will enable population surveys that reveal the origin and evolution of such sources. Proposals for other large interferometric detectors, in particular LCGT in Japan~\cite{LCGT} and AIGO in Australia~\cite{AIGO}, would add significantly to the capabilities of the current detector network. Prospects are good for detecting low-frequency signals with pulsar timing arrays on about the same time scale~\cite{PulsarTiming}, while the space-based detectors are due to be launched some years later to open up the intermediate frequency band. Conceptual designs for ``third-generation'' ground-based detectors such as the Einstein Telescope (ET)~\cite{ETweb,ETpaper} are now being proposed with the goal of improving over the sensitivities of the ``Advanced'' detectors by another order of magnitude. A comprehensive discussion of all the possible astrophysics that can be addressed is beyond the scope of this article (but see the Living Review by Sathyaprakash and Schutz~\cite{SathyaSchutz}, for example). Instead, to illustrate some of the key issues, let us look into a crystal ball (see figure~\ref{fig:CrystalBall}) \begin{figure}[bt] \begin{center} \includegraphics[width=6cm]{figure3.eps} \caption{Crystal ball for predicting the results of future gravitational-wave searches. (Glass art by Henry Summa; photo by J G Shawhan.)} \label{fig:CrystalBall} \end{center} \end{figure} and make some predictions for what the future {\em might} hold. \subsection{A wild guess at the future} Near the end of the S6/VSR2 science run, LIGO and Virgo will record a fairly significant inspiral event candidate, with a strength corresponding to a false alarm rate of 1 per 160 years. The best-match template will have masses of $8.2\,M_\odot$ and $1.46\,M_\odot$, representing a black hole and a neutron star. The all-sky burst search will also identify this as a candidate, but not the strongest one in that search. The reconstructed sky position will be consistent with three galaxies within 50~Mpc. Prompt follow-up imaging with a robotic telescope will capture a weak, fading optical transient in an elliptical galaxy within the favored sky region. After careful internal review and debate, the LIGO and Virgo collaborations will publish the complete diagnosis of the candidate, calling it ``cautious evidence for a gravitational-wave signal''. By 2015, pulsar timing array analyses will have produced greatly improved upper limits on supermassive black hole binary mergers and stochastic background processes, but no detections yet. In the Spring of 2015, Advanced LIGO and Advanced Virgo will have been mostly commissioned and (I speculate) will begin an 8-month science run while still a factor of $\sim$2 away from their design sensitivities. The GEO-HF detector may continue to operate for part of that period. The run will yield two black hole--neutron star inspiral candidates, with an expected background of $0.02$, and also two binary neutron star inspiral candidates, with a background of $0.03$. One of the binary neutron star candidates will have a clear radio afterglow in prompt follow-up observations with radio telescopes. These event candidates will be published together as the first clear detection of gravitational waves. Analysis of these events will also place strong limits on ``extra'' GW polarization states beyond the two predicted by general relativity. All-sky searches for burst and periodic GW signals using the same data will yield candidates that look promising but are not significant enough to claim as detections. Greatly improved upper limits will be published on GW emission from known pulsars and on a stochastic background. After further commissioning, Advanced LIGO and Advanced Virgo will resume running at full sensitivity, joined the following year by LCGT and AIGO. I imagine that analysis of two years of data will yield the following: \begin{itemize} \item 15 binary neutron star candidates with a background of $2.2$. Two of the candidates will correspond to short-hard GRBs, one of which is localized in a galaxy with a measured redshift of $0.07$. Based on this information, the emitted GW energy will turn out to be consistent with the theoretical prediction. \item 18 black hole--neutron star candidates with a background of $4.1$. Two of these will correspond to GRBs, one also with a high-energy neutrino. \item Comparison of GW inspiral times with GRB times will confirm that GWs travel at the speed of light. \item 6 binary black hole candidates with a background of $1.7$. The masses and spins of the candidates will be inferred, giving a preliminary look at their distributions. \item 4 burst candidates with a background of 0.15. One of them, with central frequency 310~Hz, will correspond to a weak long GRB with no measured redshift. \item A periodic GW signal will be detected from the Crab pulsar, corresponding to $0.12$\% of the total spin-down energy. This result will be used to constrain models of neutron star formation in supernovae. \item Periodic GW signals will also be detected from Sco X-1 (using data collected during a 3-month period with the detectors in a narrow-band configuration) and from 5 unseen neutron stars. \item Stochastic GW searches will detect signals from two low-mass X-ray binaries (LMXBs) besides Sco X-1, and will place much stricter limits on cosmic string models. \end{itemize} Around the same time, pulsar timing analyses will detect GWs from supermassive black hole binary systems in two galaxies, and will rule out another large area of the parameter space for cosmic string models. Gravitational-wave astronomy will thus be in full swing by the time that LISA and DECIGO are launched and open up new frequencies for GW observations. \subsection{Notes on this exercise} It may have been indulgent to speculate so specifically about what the future may bring, and the details are obviously fictional. However, I think the scenario above is actually fairly conservative and illustrates several of the scientific findings that can be derived from the observations. It also reflects many of the issues the GW community will have to deal with, such as borderline-significant event candidates, samples of event candidates with non-negligible backgrounds, and the role of information from electromagnetic observations. \section{The impact of detections on the field of gravitational-wave science} The transition envisioned above, from always setting upper limits to being able to claim some detections, will call for some changes in strategy to make optimal use of the detectors and of the data for science results. In this section we discuss a few such areas. \subsection{Tuning choices for advanced detectors} Interferometric detectors may be operated in different ways in order to optimize the noise characteristics according to scientific priorities. For instance, the Advanced LIGO and Advanced Virgo detectors are designed to be limited by quantum noise at low and high frequencies; by reducing the laser intensity, one can reduce the radiation pressure noise at low frequency at the cost of increasing the shot noise at high frequency. Interferometer configurations with signal recycling, such as Advanced LIGO and Advanced Virgo, allow additional tuning options through changing the reflectivity and (microscopic) detuning phase shift of the signal recycling mirror. Optimal tunings have been considered for individual signal types as well as some combinations~\cite{AdLIGOtuning}. Of course, an interferometer can only operate in one mode at a time. Detection of one or more GW signals may motivate re-tuning the interferometer to focus on a certain class of signals, either temporarily or for the rest of the run. It may also be useful to tune different interferometers differently, {\it e.g.}\ one of the Advanced LIGO Hanford interferometers could be optimized for low frequency while the other is optimized for higher frequency. \subsection{Support for additional detectors} The first direct detection of a GW signal will erase any lingering doubts about whether GW detectors really work, and will bring a new focus to the science which can be done with GW observations. That should make a stronger case for additional detectors on the ground ({\it e.g.}\ LCGT, AIGO, ET) as well as boosting support for detectors in space (LISA, DECIGO). Furthermore, the designs of new detectors may be influenced by the view of what measurements are most important based on what has been detected so far. \subsection{Changes in philosophy} Currently there is a big emphasis in the GW detection community on achieving near-perfect certainty in the first GW signal detection. Typically this results in raising the signal strength threshold so that the false alarm rate is extremely low, but that also reduces the sensitivity of the search. However, having certified one or more events as genuine increases our belief in other candidates of the same type. Thus we can choose to relax the signal strength threshold to include more candidates in a search, even if doing so also includes more background---the benefit from having more real signals in the sample to study may outweigh the negative effects of the additional background. \section{Summary} Many of the gravitational-wave searches that have been performed in recent years have provided useful astrophysical information despite yielding no confirmed GW signal candidates. Thus, one can say that we are already doing gravitational-wave astronomy. Actual detections, when they finally start coming, will enable us to address a much wider range of astrophysics questions. And here is what will be exciting: {\em Everything}. \ack I would like to thank the organizers of the Eighth Edoardo Amaldi Conference on Gravitational Waves for giving me this opportunity to review and interpret the current state of gravitational-wave astronomy. Of course, my colleagues in the gravitational-wave community are responsible for the observational results themselves, and my views of the astrophysics and of the field have been shaped by discussions with many of them---too many to thank individually. I was particularly inspired by a 2007 seminar by Ben Owen entitled ``Why LIGO results are already interesting''. I gratefully acknowledge the support of the National Science Foundation through grant PHY-0757957. This article has been assigned LIGO document number P0900289-v4. \section*{References}
\section{Introduction} The interaction of a quantum system with an environment is generally seen as source of decoherence, leading to the loss of quantum properties and the appearance of the classical world \cite{Breuer}. In fact, due to the creation of correlation with the environment, the evolution of the quantum system is not unitary any more. If the system is prepared in an entangled state, such entanglement is very likely to be lost \cite{YuScience}. However, environments can also create correlations among quantum systems. For example, the interaction of a bipartite system with a common reservoir creates correlations between the parts \cite{Benatti}, irrespective of the existence of a direct coupling between them \cite{Ficek,FicekESB}. In a shared environment highly entangled long-living states, or even sub-radiant decoherence free states can exist \cite{Tanas}. Many studies have demonstrated the existence of non-zero stationary entanglement in systems of two qubits or two harmonic oscillators sharing the same reservoir \cite{Paz}. The ability to create correlations persists if the common reservoir has non zero temperature \cite{Braun}, and is increased by memory effects in non-Markovian reservoirs \cite{Maniscalco,Francica,Mazzola}. In the last decade many efforts have been done to understand the dynamics of entanglement and quantum correlations in systems which are good candidates for applications in quantum information theory and technology. However, not all the protocols in quantum information theory rely on entanglement. On the contrary, in many cases entanglement is simply not enough and non-local properties, expressed by the violation of a Bell inequality, are required. Here I study the non-local properties, namely the violation of the CHSH inequality, of a system of two qubits interacting with a common non-Markovian reservoir. I consider a factorized state of the qubits with one excitation, and study, as a function of time and of the qubit-cavity coupling, the sudden violation of CHSH inequality. The conditions for which the violation is maximum are identified. Afterwards I add another dynamical ingredient and study how the dynamics of the CHSH violation gets modified when the spontaneous emission of the qubits is included. \section{The model} The system under investigation comprises two qubits interacting with a common zero-temperature bosonic reservoir. The Hamiltonian of the system in rotating-wave approximation, and in units of $\hbar$, is $H=H_{0}+H_{\mathrm{int}}$, \begin{equation} H_{0}=\omega_{0}(\sigma_{+}^{(1)}\sigma_{-}^{(1)}+\sigma_{+}^{(2)}\sigma_{-}^{(2)}) +\sum_{k}\omega_{k}a_{k}^{\dag}a_{k},\label{H0bare} \end{equation} \begin{equation}\label{Hintbare} H_{\mathrm{int}}=(\alpha_{1}\sigma_{+}^{(1)}+\alpha_{2}\sigma_{+}^{(2)})\sum_{k}g_{k}a_{k}+\mathrm{h.c. }, \end{equation} where $\sigma_{\pm}^{(1)}$ and $\sigma_{\pm}^{(2)}$ are the Pauli raising and lowering operators for qubit 1 and 2 respectively, $\omega_{0}$ is the Bohr frequency of the two identical qubits, $\alpha_{1}$ and $\alpha_{2}$ are dimensionless environment-qubit coupling constants, $a_{k}$ and $a_{k}^{\dagger}$, $\omega_{k}$ and $g_{k}$ are the annihilation and creation operators, the frequency and the coupling constant of the field mode $k$, respectively. In the following I assume that the two qubits interact resonantly with a lossy cavity, so I choose a Lorentzian spectral distribution to describe the properties of the environment, \begin{equation} J(\omega)=\frac{W^2}{\pi}\frac{\lambda}{(\omega-\omega_{0})^2+\lambda^2}, \end{equation} where $\lambda$ is the width of the spectral distribution describing the cavity losses, and $W$ in the limit of ideal cavity (when $\lambda\rightarrow0$) is proportional to the vacuum Rabi frequency $R$ through $W=R/\alpha_{T}$ with $\alpha_{T}=(\alpha_{1}^2+\alpha_{2}^2)^{1/2}$ the collective coupling constant. The dynamics of two qubits interacting with a common Lorentzian-structured reservoir has been studied in Ref. \cite{Maniscalco} for the case of one excitation, and in Ref. \cite{Mazzola} for a generic state of two identically coupled qubits. Since here I consider the dynamics of a factorized state with one excitation of the form $\ket{\Psi(t)}=c_{1}(t)\ket{10}\ket{0}_{E}+c_{2}(t)\ket{01}\ket{0}_{E}+\sum_{k}c_{k}(t)\ket{00}\ket{1_{k}}_{E}$ I am going to use the model of Ref. \cite{Maniscalco}. There the dynamics of the qubits is expressed in the basis of super-radiant and sub-radiant states \begin{equation}\label{Psim} \ket{\psi_{-}}=r_{2}\ket{10}-r_{1}\ket{01}, \end{equation} \begin{equation}\label{Psip} \ket{\psi_{+}}=r_{1}\ket{10}+r_{2}\ket{01}, \end{equation} where the relative coupling strengths $r_{1}=\alpha_{1}/\alpha_{T}$ and $r_{2}=\alpha_{2}/\alpha_{T}$ have been introduced ($r_{1}^2+r_{2}^2=1$). Considering the dynamics in such a basis is particularly convenient since the sub-radiante state does not evolve in time. Thus the evolution of the amplitudes of the first and second qubits is just \begin{equation} c_{1}(t)=r_{2}\beta_{-}+r_{1}\beta_{+}E(t), \end{equation} \begin{equation} c_{2}(t)=-r_{1}\beta_{-}+r_{2}\beta_{+}E(t), \end{equation} with $\beta_{\pm}=\bra{\psi_{\pm}}\psi_{0}\rangle$ and \begin{equation} E(t)=e^{-\lambda t/2}[\cosh(\Omega t/2)+\frac{\lambda}{\Omega}\sinh(\Omega t/2)], \end{equation} where $\Omega=\sqrt{\lambda^2-4R^2}$. In order to evaluate the time evolution of the violation of the CHSH Bell inequality I use the expression of Ref. \cite{BellomoNLE}, which allows one to express the maximum of the Bell function (by an appropriate choice of angles) as function of the two-qubit density matrix elements. In the $\{\ket{11},\ket{10},\ket{01},\ket{00}\}$ basis of the qubits, the maximum of the Bell function reads as follows \begin{equation} B=2 \max_{i,j}\{u_{i}+u_{j}\}^{1/2} \end{equation} with \begin{eqnarray} u_{1}=(\rho_{11}+\rho_{44}-\rho_{22}-\rho_{33})^2,\nonumber\\ u_{2}=4(|\rho_{23}|^2+|\rho_{14}|^2),\\ u_{3}=4(|\rho_{23}|^2-|\rho_{14}|^2).\nonumber \end{eqnarray} where $\rho_{22}=|c_{1}(t)|^2$, $\rho_{33}=|c_{2}(t)|^2$, $\rho_{23}=c_{1}(t)c_{2}^{*}(t)$ and $\rho_{44}=1-|c_{1}(t)|^2-|c_{2}(t)|^2$. For the particular initial state considered I have $\rho_{11}=0$ and $\rho_{14}=0$. \section{CHSH violation dynamics} Here I investigate the CHSH inequality violation for two qubits prepared initially in the state \begin{equation}\label{Psi0} \ket{\psi(0)}=\ket{10}. \end{equation} I consider a cavity characterized by a high but experimentally feasible quality factor (the width is $\lambda=10^{-1}\ \gamma_0$, with $\gamma_{0}$ the Markovian decay rate of the qubits), and strong coupling conditions. I set the parameter $S=R/\lambda=10$, indicating the strength of the coupling \cite{Maniscalco}. In figure \ref{fig:CHSH} I plot the CHSH violation as a function of scaled time ($\tau=\lambda t$) and of the relative coupling strength $r_{1}$. During the evolution the system passes through highly entangled states, strongly violating the CHSH inequality. Cycles of birth and death of non-locality follow one after the other, till for long times the inequality is finally satisfied. The sudden appearance of the violation is consequence of the shared reservoir, providing an indirect coupling between the qubits. Such a reservoir-mediated interaction plays certainly a role also in the revivals of non-locality, which are also caused by the memory effects of the non-Markovian reservoir. I notice that sudden violation of the CHSH inequality appears up to a certain value of the relative strength of the coupling. In particular for the chosen coupling conditions no violation occurs when $r_{1}>0.6$. For stronger coupling conditions the region of violation in the $r_{1}$ space of parameters widens, nevertheless it never reaches $1/\sqrt{2}$. In fact for $r_{1}=r_{2}=1/\sqrt{2}$ the qubits are symmetrically coupled to the cavity, and entanglement and non-local properties in general cannot be created out of the factorized state in Eq. \eref{Psi0}. \begin{figure}[!] \begin{center} \includegraphics[width=12cm]{CHSH.eps} \caption{\label{fig:VNCBell2}(Color online) Dynamics of the CHSH inequality violation in a common Lorentzian structured reservoir as a function of scaled time and relative coupling strength for two atoms prepared in a factorized state $\Psi_{0}=\ket{10}$. No spontaneous emission by the qubits is included.}\label{fig:CHSH} \end{center} \end{figure} \section{Including spontaneous emission} In this section I study how the dynamical violation of the CHSH inequality is modified when including spontaneous emission for the two qubits. In Ref. \cite{MazCross} I have studied the entanglement time evolution of two entangled qubits interacting with the same Lorentzian structured reservoir (leaky cavity) and emitting independently outside the cavity due to spontaneous emission. By means of Eqs. (1) and (2) of Ref. \cite{MazCross} I have solved the dynamics in the case of qubits having the same transition frequency, equally and resonantly coupled with the cavity. Here I generalize that master equation to include the case of different couplings between the qubits and the cavity. So the master equation describing the dynamics of our system (with some renaming of parameters) becomes: \begin{eqnarray}\label{ME} \frac{\partial\tilde{\rho}}{\partial t}=-\imath[H,\tilde{\rho}] -\lambda(a^{\dag}a\tilde{\rho}+\tilde{\rho}a^{\dag}a-2a\tilde{\rho}a^{\dag})\nonumber\\ -\frac{\gamma_{1}}{2}(\sigma_{+}^{(1)}\sigma_{-}^{(1)}\tilde{\rho}+\tilde{\rho}\sigma_{+}^{(1)}\sigma_{-}^{(1)}-2\sigma_{-}^{(1)}\tilde{\rho}\sigma_{+}^{(1)})\nonumber\\ -\frac{\gamma_{2}}{2}(\sigma_{+}^{(2)}\sigma_{-}^{(2)}\tilde{\rho}+\tilde{\rho}\sigma_{+}^{(2)}\sigma_{-}^{(2)}-2\sigma_{-}^{(2)}\tilde{\rho}\sigma_{+}^{(2)}), \end{eqnarray} with \begin{equation} H=\alpha_{T}W[(r_{1}\sigma_{+}^{(1)}+r_{2}\sigma_{+}^{(2)})a+\mathrm{h.c. }]. \end{equation} where $\sigma_{\pm}^{(1)}$ and $\sigma_{\pm}^{(2)}$, $\alpha_{t}$, $r_{1}$ and $r_{2}$, $\lambda$ and $W$ have been defined above, while $a$ and $a^{\dagger}$ are the annihilation and creation operators for the cavity mode and $\gamma_{1/2}$ are the spontaneous emission rates for the two qubits. For the sake of simplicity I assume $\gamma_{1}=\gamma_{2}=\gamma_{S}$. As in the previous section I consider the sudden appearance of CHSH violation for two qubits prepared in the state $\ket{\psi(0)}=\ket{10}$. I study the dynamics for the same strong coupling conditions $S=R/\lambda=10$, in particular for the same width of the spectral distribution, and vacuum Rabi frequency. The two-qubit spontaneous emission decay rates are set equal to $\gamma_{S}/\gamma_{0}=1/50$, therefore they are 50 times smaller than the vacuum Rabi frequency $R$. In figure \ref{fig:CHSHspont} I show the CHSH violation as a function of time and of the relative coupling parameter $r_{1}$. The effect of spontaneous emission is apparent: the long series of sharp revivals is dramatically damped out. For the chosen spontaneous emission parameter only one small revival is present. The amount of violation of the CHSH inequality reduces with increasing decay rate, and beyond a certain threshold value the CHSH inequality is always satisfied. For the coupling conditions chosen such a threshold parameter is $\gamma_{S}=1/9 \ \gamma_{0}$. \begin{figure}[!] \begin{center} \includegraphics[width=12cm]{CHSHspont.eps} \caption{\label{fig:VNCBell2}(Color online) Dynamics of the CHSH inequality violation in a common Lorentzian structured reservoir as a function of scaled time and relative coupling strength for two atoms prepared in a factorized state $\Psi_{0}=\ket{10}$. Here spontaneous emission is included and the spontaneous emission decay rate of the qubits are equal to $(1/50) \gamma_{0}$, i.e., 50 times smaller than the vacuum Rabi frequency $R=\alpha_{T}W$.}\label{fig:CHSHspont} \end{center} \end{figure} \section{Conclusion} I have investigated the time evolution of the violation of the CHSH inequality for a system of two qubits interacting with a common Lorentzian structured reservoir. The reservoir-induced correlations between the qubits drive the evolution of an initial factorized state as $\ket{\psi(0)}=\ket{10}$ towards highly entangled states, strongly violating the CHSH inequality. After cycles of birth and death of non-locality, progressively damped out, the CHSH inequality becomes always satisfied. As a second step I consider a more realistic system in which spontaneous emission of the two qubits is taken into account. The time evolution of the CHSH violation well describes the dramatic changes that the introduction of spontaneous emission in this model brings to the dynamics of the system. Birth of non-local properties appears only if the spontaneous emission decay rate is below a certain threshold, even in that case oscillations are strongly damped and the inequality becomes permanently satisfied much earlier. The dynamics of the CHSH violation was studied for a system of two non interacting initially entangled qubits in two independent non-Markovian reservoirs \cite{BellomoNLE,Bellomo}. There, death and revivals of the violation have been observed due to the memory effects of the reservoirs. In that case however, correlations between the qubits cannot be created starting from a factorized state, as a consequence an initially separable state cannot evolve into an entangled one, and possibly violating the CHSH inequality. \ack I would like to thank S. Maniscalco and J. Piilo for enlightening discussions, G. Compagno and his group in Palermo for the kind hospitality and useful suggestions. Finally I thank M. Ehrnrooth Foundation for financial support. \section*{References}
\section{Introduction} In addition to the mass, metallicity is one of the few progenitor attributes that can leave an imprint on the observational properties of SNIa by affecting the synthesized mass of $^{56}$Ni, with important consequences for their use as cosmological standard candles. Up to now, attempts to measure $Z$ directly from supernova observations have been scarce and their results uncertain \citep{len00,tau08}. Measuring $Z$ from the X-ray emission of supernova remnants is a promising alternative but as yet has been only applied to a single supernova \citep{bad08a}. An alternative venue is to estimate the supernova metallicity as the mean $Z$ of its environment \citep{bad09}. \citet{ham00} looked for galactic age or metal content correlations with SNIa luminosity, but their results were ambiguous. \citet{ell08} looked for systematic trends of SNIa UV spectra with metallicity of the host galaxy, and found that the spectral variations were much larger than predicted by theoretical models. \citet{coo09}, using data from the Sloan Digital Sky Survey and Supernova Survey concluded that prompt SNIa are more luminous in metal-poor systems. Recently, \citet[hereafter G08]{gal08} and \citet[hereafter H09]{how09}, using different methodologies to estimate the metallicity of SNIa hosts, arrived to opposite conclusions with respect to the dependence of supernova luminosity on $Z$. There is a long history of numerical simulations of SNIa aimed at predicting the impact of metallicity and explosive neutronization on their yields \cite[e.g.][]{bra92,brach00,tra05,bad08a}. \citet[hereafter DHS01]{dom01} found that the offset in the calibration of supernova magnitudes vs light curve (LC) widths is not monotonic in $Z$ and remains smaller than $0.07^\mathrm{m}$ for $Z\leq0.02$. \citet{kas09} concluded that the width-luminosity relationship depends weakly on the metallicity of the progenitor. From an analytical point of view, \citet[hereafter TBT03]{tim03} using arguments from basic nuclear physics predicted a linear relationship between $M(^{56}\mathrm{Ni})$ and $Z$. The conclusions of TBT03 relied on two main assumptions: first, that most of the $^{56}$Ni is synthesized in material that burns fully to NSE and, second, that a fiducial SNIa produces a mass $M_\mathrm{Fe}^{\eta_0}\sim0.6$~M$_\odot$ of Fe-group nuclei whose $\eta$ is not modified during the explosion. \citet{pb08} and \citet{ch08}, based on the same assumptions as TBT03, extended their analysis taking into account the neutronization produced during the simmering phase. In this paper, we show that the first assumption of TBT03 does not hold for most SNIa. Indeed, for a SNIa that produces $M_\mathrm{Fe}^{\eta_0}\sim0.6$~M$_\odot$ the fraction of $^{56}$Ni synthesized out of NSE exceeds $\sim30\%$. With respect to the second assumption, \citet[hereafter M07]{maz07} showed, based on observational results, that the mass of Fe-group nuclei ejected by SNIa spans the range from 0.4 to 1.1~M$_\odot$. This range cannot be accounted for by metallicity variation within reasonable values. Accordingly, our working hypothesis is that the yield of $^{56}$Ni in SNIa is governed by a primary parameter different from $Z$. In our one-dimensional models the primary parameter is the DDT density, $\rho_\mathrm{DDT}$, although in nature it may be something else such as the expansion rate during the deflagration phase. The initial metallicity is a secondary factor that can give rise to scatter in the value of $M(^{56}\mathrm{Ni})$ either directly ({\sl linear scenario}), by affecting the chemical composition of the ejecta for a given value of the primary parameter, or indirectly ({\sl non-linear scenario}), by modifying the primary parameter itself. The understanding of which one of these two characters is actually being played by $Z$ is of paramount importance. \section{The effect of metallicity on the yield of $^ {56}$Ni} We explore the sensitivity of $M(^{56}\mathrm{Ni})$ to the progenitor metallicity, starting from Pre-Main Sequence models of masses, $M_0$, in the range $2-7$~M$_\odot$ and metallicities, $Z$, from $10^{-5}$ to 0.10, as given in the first column of Table~\ref{tab1}. The initial mass fractions of all the isotopes with $A\gtrsim6$ have been fixed in solar proportion, according to \citet{lod03}; consequently, we adopt for the solar metallicity the value $Z_\odot =0.014$. Each presupernova model has been evolved from the Pre-Main Sequence to the Thermal Pulse (TP) AGB phase, in order to determine the mass, $M_\mathrm{core}$, and chemical structure of the C-O core left behind. Afterwards, an envelope of the appropriate size to reach the Chandrasekhar mass, $M_\mathrm{Ch}$, has been added on top of the C-O cores, and these structures have been fed as initial models to a supernova hydrocode. Finally, the explosive nucleosynthesis has been obtained with a post-processing nucleosynthetic code. The hydrostatic evolution has been computed by means of the FRANEC code \citep{chi98}. With respect to the calculations of DHS01, the code has been updated in the input physics. For the purposes of the present paper, the most important changes concern the $^{12}\mathrm{C}\left(\alpha,\gamma\right)^{16}\mathrm{O}$ reaction rate, which is calculated according to \citet{kun02} instead of \citet{cf85}, and the treatment of convective mixing during the late part of the core-He burning \citep{str03}. The presupernova model is a Chandrasekhar mass WD built in hydrostatic equilibrium with a central density $\rho_\mathrm{c}=3\times10^9$~g~cm$^{-3}$. The composition of the envelope of mass $M_\mathrm{Ch}-M_\mathrm{core}$ is the same as that of the outermost shell of the C-O core. Thus, instead of assuming C/O=1, as in DHS01, we adopt the C/O ratio obtained as a result of He-shell burning during the AGB phase. The effect of changing $\rho_\mathrm{c}$ and the composition of the envelope has been tested in several models, as explained later. We leave aside other eventual complexities of pre-supernova physics like rotation \citep{pie03,yoo04}. The internal composition of the WD is eventually modified during the simmering phase, due to the combined effects of convective mixing, carbon burning and electron captures. The first two phenomena affect the carbon abundance within the core, while the latter leads to an increase of $\eta$. The average (within the WD) carbon consumption and neutron excess increase during the simmering phase are $\Delta Y(^{12}\mathrm{C})\approx-1.66\times10^{-3}$~mol~g$^{-1}$ and $\Delta\eta =-\frac{2}{3}\Delta Y(^{12}\mathrm{C})$ \citep{ch08}. We assume that convective mixing is limited to the C-O core, which implies that the change in the neutron excess {\sl within the core} is $\Delta\eta\approx 1.11\times10^{-3}M_\mathrm{Ch}/M_\mathrm{core}$. We have also exploded several models disregarding the simmering phase, to which we will refer in the following as {\sl stratified models}. The supernova hydrodynamics code we have used is the same as in \citet{bad03}. As in DHS01, the present models are based on the delayed-detonation paradigm \citep{kho91}. To address the {\sl linear scenario} we take $\rho_\mathrm{DDT}$ independent of $Z$. In this case, $\rho_\mathrm{DDT}=3\times10^7$~g~cm$^{-3}$, although simulations with $\rho_\mathrm{DDT}$ in the range $1-3\times10^7$~g~cm$^{-3}$ are also reported. For the {\sl non-linear scenario} we have adopted the criterion that a DDT is induced when the laminar flame thickness, $\delta_\mathrm{lam}$, becomes of the order of the turbulent Gibson length $l_\mathrm{G}$ \citep{rop07}, with the flame properties (velocity and width) depending on the abundances of $^{12}$C \citep[eq. 22 in][]{woo07b} and $^{22}$Ne \citep{cha07}, and hence on $Z$ and $\eta$. \citet{tow09} concluded from 2D simulations of SNIa that the metallicity does not affect the dynamics of the explosion, and so the turbulence intensity is independent of $Z$. Thus, for a given turbulent intensity a change in $Z$ can be compensated by a change in $\rho_\mathrm{DDT}$ in order to recover the condition $\delta_\mathrm{lam}/l_\mathrm{G}\approx1$ \citep[see the discussion in][]{cha07}. In this scenario we have scaled the transition density as a function of the local chemical composition as follows: \begin{equation} \rho_\mathrm{DDT}\propto X(^{12}\mathrm{C})^{-1.3}\left(1+129\eta\right)^{-0.6}\,. \label{eq2.5} \end{equation} \noindent In order to introduce an $\eta$ dependence in the above expression we have assumed, for simplicity, that the bulk of neutronized isotopes synthesized during the simmering phase accelerates the carbon consumption rate the same way $^{22}$Ne does. \subsection{Presupernova evolution} The results of the hydrostatic evolution of our presupernova models are shown in Table~\ref{tab1}. For each $M_0$ and $Z$ we give: $M_\mathrm{core}$, the central abundance of $^{12}$C and $\eta$ in stratified models, $X_\mathrm{c}(^{12}\mathrm{C})$ and $\eta_\mathrm{c}$, and the same quantities in the models accounting for the simmering phase, $X_\mathrm{sim}(^{12}\mathrm{C})$ and $\eta_\mathrm{sim}$. In comparison with DHS01, the present models span a larger range of $Z$, as DHS01 computed models with $Z\lesssim Z_\odot$. In the common range of $Z$ and $M_0$ the results are comparable, although the adopted rate of the $^{12}\mathrm{C}\left(\alpha,\gamma\right)^{16}\mathrm{O}$ reaction leads to a slightly larger carbon abundance than in DHS01. The differences in $M_\mathrm{core}$ between our models and those of DHS01 are smaller than $0.06$~M$_\odot$. The central carbon to oxygen ratio and $M_\mathrm{core}$ we obtain, and their dependencies with $Z$ and $M_0$, agree as well with \citet{ume99}. \subsection{Mass of $M(^{56}\mathrm{Ni})$ ejected} The results of the explosion simulations are summarized in Figs.~\ref{fig1} and \ref{fig2}. Figure~\ref{fig1} shows the dependence of $M(^{56}\mathrm{Ni})$ on $Z$. For the stratified models, we obtain the same range of variation of $M(^{56}\mathrm{Ni})$ with respect to $M_0$ at given $Z$ as DHS01: 0.06~M$_\odot$, although the $^{56}\mathrm{Ni}$ yields do not match with DHS01 because they used different values of $\rho_\mathrm{c}=2\times10^9$~g~cm$^{-3}$ and $\rho_\mathrm{DDT}=2.3\times10^7$~g~cm$^{-3}$. The models accounting for the simmering phase behave like the stratified models with respect to variations in $M_0$ and $Z$, although with a smaller total $M(^{56}\mathrm{Ni})$ due to electron captures during the simmering phase. The dependence of $M(^{56}\mathrm{Ni})$ on $Z$ can be approximated by a linear function: \begin{equation} M(^{56}\mathrm{Ni})\propto f(Z)=1-0.075\frac{Z}{Z_\odot}\,, \label{eq0} \end{equation} \noindent while stratified models can be approximated by: $M(^{56}\mathrm{Ni})\propto1-0.069Z/Z_\odot$, i.e. the slope of the linear function is quite insensitive to the carbon simmering phase. To explore the {\sl non-linear scenario} we have computed models accounting for the simmering phase with fixed $M_0 = 5$~M$_\odot$. Introducing a composition dependent $\rho_\mathrm{DDT}$ produces a qualitatively different result because the relationship between $M(^{56}\mathrm{Ni})$ and $Z$ is no longer linear, especially at low metallicites for which a larger $\rho_\mathrm{DDT}$ is obtained, implying a much larger $M(^{56}\mathrm{Ni})$. Our results can be fit by a polinomial law: \begin{equation} M(^{56}\mathrm{Ni})\propto f(Z)=1-0.18\frac{Z}{Z_\odot}\left(1-0.10\frac{Z}{Z_\odot}\right)\,. \label{eq1} \end{equation} Both the central density at the onset of thermal runaway and the final C/O ratio in the accreted layers have a minor effect on $M(^{56}\mathrm{Ni})$ within the explored range. TBT03 proposed a linear relationship between $M(^{56}\mathrm{Ni})$ and $Z$: $M(^{56}\mathrm{Ni})\propto1-0.057Z/Z_\odot$ (dotted line in Fig.~\ref{fig1}). In all of our present models we find a steeper slope. The reason for this discrepancy lies in the assumption by TBT03 that most of the $^{56}$Ni is synthesized in NSE. In our models a sizeable fraction of $^{56}$Ni is always synthesized during incomplete Si-burning, whose final composition has a stronger dependence on $Z$ than NSE matter. As \citet{hix96} showed, the mean neutronization of Fe-peak isotopes during incomplete Si-burning is much larger than the global neutronization of matter because neutron-rich isotopes within the Si-group are quickly photodissociated, providing free neutrons that are efficiently captured by nuclei in the Fe-peak group, favouring their neutron-rich isotopes. Figure~\ref{fig2} shows that up to $\sim60\%$ of $M_\mathrm{Fe}^{\eta_0}$ can be made out of NSE, the actual fraction depending essentially on the total mass of Fe-group elements ejected. Thus, the less $M(^{56}\mathrm{Ni})$ is synthesized, the larger fraction of it is built during incomplete Si-burning and the stronger is its dependence on $Z$. \section{Discussion} The results presented in the previous section show that the metallicity is not the primary parameter that allows to reproduce the whole observational scatter of $M(^{56}\mathrm{Ni})$, for a reasonable range of $Z$. We have also shown that a possible dependence of the primary parameter on $Z$, would lead to a non-linear relationship between $M(^{56}\mathrm{Ni})$ and $Z$, as in Eq.~\ref{eq1}. However, as we will show in the following, it would be possible to unravel the way $M(^{56}\mathrm{Ni})$ depends on $Z$ by means of future accurate measurements of SNIa properties. We start analysing the amount of the scatter induced by the dependence of $M(^{56}\mathrm{Ni})$ on $Z$ given by Eq.~\ref{eq1}. For simplicity we follow the procedure of M07 to estimate the supernova luminosity and LC width. The peak bolometric luminosity, $L$, is determined directly by the mass of $^{56}$Ni synthesized (in the following, all masses are in $M_\odot$ and energies are in $10^{51}$~ergs): \begin{equation} L\left[M(^{56}\mathrm{Ni})\right]=2\times10^{43}M(^{56}\mathrm{Ni})~\mathrm{erg}~\mathrm{s}^{-1}\,, \label{eq1.5} \end{equation} \noindent while the bolometric LC width, $\tau$, is determined by the kinetic energy, $E_\mathrm{k}$, and the opacity, $\kappa$: $\tau\propto\kappa^{1/2}E_\mathrm{k}^{-1/4}$. The kinetic energy is given by the difference of the WD initial binding energy, $\mathrm{|BE|}$, and the nuclear energy released, the latter being related to the final chemical composition of the ejecta: $E_\mathrm{k} \approx1.56M(^{56}\mathrm{Ni})+1.74\left[M_\mathrm{Fe}-M(^{56}\mathrm{Ni})\right]+1.24M_\mathrm{IME }-\mathrm{|BE|}$ , where $M_\mathrm{Fe}$ is the total mass of Fe-group nuclei and $M_\mathrm{IME}$ is the mass of intermediate-mass elements (IME). The opacity is provided mainly by Fe-group nuclei and IMEs: $\kappa\propto M_\mathrm{Fe}+0.1M_\mathrm{IME}$. We have taken $\mathrm{|BE|}=0.46$, which is a good approximation given the small variation of binding energy with initial central density: $\mathrm{|BE|}$ is in the range $0.44-0.47$ for $\rho_\mathrm{c}=2-4\times10^9$~g~cm$^{-3}$. To reduce the number of free parameters we further link $M_\mathrm{IME}$ to $M_\mathrm{Fe}$ imposing that the ejected mass is the Chandrasekhar mass ($M_\mathrm{Ch}\approx1.38~\mathrm{M}_\odot$ in our models), and that the amount of unburned C+O scales as $M_\mathrm{CO}\approx0.3M_\mathrm{IME}^2$, as deduced from our models. Thus, $M_\mathrm{Fe}+M_\mathrm{IME}+0.3M_\mathrm{IME}^2=M_\mathrm{Ch}$. Furthermore, the mass of $^{56}$Ni is linked to the mass of Fe-group nuclei by $M(^{56}\mathrm{Ni})=M_\mathrm{Fe}^{\eta_0}\times f(Z)=\left(M_\mathrm{Fe}-M_\mathrm{ec}\right)\times f(Z)$, where $f(Z)$ is given by Eq.~\ref{eq1} or a similar function, and $M_\mathrm{ec}$ is the mass of the neutron-rich Fe-group core (due to electron captures during the explosion). We have taken $M_\mathrm{ec}\approx0.14~\mathrm{M}_\odot$, which is representative of the range of masses obtained in our models: $0.10-0.16~\mathrm{M}_\odot$ for $\rho_\mathrm{c}=2-4\times10^9$~g~cm$^{-3}$. Finally, to compare with observed values a scale factor of 24.4 is applied to the value of $\tau$ thus obtained, as in M07. Putting all these together, we obtain the following expression for the bolometric LC width (in days) as a function of $M(^{56}\mathrm{Ni})$ and $Z$: \begin{equation} \tau\left[M(^{56}\mathrm{Ni}),Z\right] = 21.9\frac {\left\lbrace\displaystyle{\frac{M(^{56}\mathrm{Ni})}{f(Z)}}-0.027+0.263\sqrt{1-0.482\displaystyle{ \frac{M(^{56}\mathrm{Ni})} {f(Z)}}}\right\rbrace^{1/2}} {\left\lbrace\left(\displaystyle{\frac{1.115}{f(Z)}}-0.115\right)M(^{56}\mathrm{Ni})-1.46+2.09\sqrt { 1-0.482\displaystyle{\frac{M(^{56}\mathrm{Ni})}{f(Z)}}}\right\rbrace^{1/4}}\,. \label{eq2} \end{equation} \noindent The relationship between $L$ and $\tau$ derived from Eqs.~\ref{eq1}, \ref{eq1.5} and \ref{eq2} is displayed in Fig.~\ref{fig3} for three different metallicities along with observational data. There are also represented the relationships obtained by substituting Eq.~\ref{eq1} by the $M(^{56}\mathrm{Ni})$ vs $Z$ dependences proposed by TBT03 and Eq.~5 in H09. Our Eq.~\ref{eq1} gives a wider range of $M(^{56}\mathrm{Ni})$, which accounts better for the scatter of the observational data. Indeed, if real SNIa follow Eq.~\ref{eq1}, deriving supernova luminosities from $Z$-uncorrected LC shapes might lead to systematic errors of up to 0.5~magnitudes. To estimate the bearing that the metallicity dependence of $M(^{56}\mathrm{Ni})$ can have on cosmological studies that use a large observational sample of supernovae, we have generated a virtual population of 200 SNIa that has been analyzed following the same methodology as G08 and H09. Each virtual supernova has been randomly assigned a progenitor metallicity, from a uniform distribution of $\log(Z)$ between $Z_\mathrm{min}=0.1Z_\odot$ and $Z_\mathrm{max}=3Z_\odot$, and an $M_\mathrm{Fe}$, uniformly distributed in the range from 0.31 to 1.15~$\mathrm{M}_\odot$. The minimum and maximum $M(^{56}\mathrm{Ni})$ thus obtained (computed with Eq.~\ref{eq1} and $M_\mathrm{ec}=0.14~\mathrm{M}_\odot$) are $0.1$ and $1~\mathrm{M}_\odot$, and the bolometric LC width, $\tau$, lies in the range $15-24$~days. A $Z$-uncorrected mass of $^{56}$Ni, $M^{\odot}_{56}$, has then been obtained as the value of $M(^{56}\mathrm{Ni})$ that would give the same $\tau$ if $Z=Z_\odot$. The $M^{\odot}_{56}$ so computed gives an idea of the effect of fitting an observed SNIa LC with a template that takes no account of the supernova metallicity. From Eq.~\ref{eq1.5}, we estimate the Hubble Residual, HR, of each virtual SNIa at: $\mathrm{HR}=2.5\log\left(M^{\odot}_{56}/M(^{56}\mathrm{Ni})\right)$. As a final step we have added gaussian noise with $\sigma=0.1$ to both HR and $\log(Z)$, to simulate the effect of observational uncertainties. A linear relationship $\mathrm{HR}=\alpha+\beta\log(Z)$ has then been fit to the noisy virtual data by the least-squares technique, as in G08 and H09. Figure~\ref{fig4} shows the results for 10,000 realizations of the noisy virtual dataset. The histogram gives the number counts of the slope $\beta$ in the 10,000 realizations. The whole process has been repeated by using Eq.~\ref{eq0} (i.e. the {\sl linear scenario}) to represent the dependence of $M(^{56}\mathrm{Ni})$ on $Z$ and the results are also shown in Fig.~\ref{fig4}. From the Figure it is clear that, for a large enough set of SNIa whose luminosity and metallicity are measured with small enough errors, it is possible to discriminate between the {\sl linear} and {\sl non-linear scenarios}. In our numerical experiment, the mean value of $\beta$ is 0.13 in the first case and 0.26 in the second case, both with a standard deviationof 0.02. Figure~\ref{fig4} shows also the observational results obtained by G08, who approximated the metallicities of the SNIa in their sample by the $Z$ of the host galaxy, obtained from an empirical galactic mass-metallicity relationship. The striking match between our results based on the {\sl non-linear scenario} and those of G08 must be viewed with caution in view of the observational uncertainties involved in measuring supernova metallicities and the limitations of our models (i.e. the assumption of spherical symmetry). Recently, using a different method of determination of the SNIa metallicity, H09 arrived to a result opposite to that of G08, i.e. they found that HR is uncorrelated with $Z$, leading to a distribution centered around $\beta\approx0$. Thus, until such discrepancies are resolved it is not possible to draw any firm conclusion about the metallicity effect on SNIa luminosity. However, it is worth stressing that the simultaneous measurement of supernova luminosity and metallicity for a large SNIa set would strongly constrain the physics of the deflagration-to-detonation transition in thermonuclear supernovae, one of the key standing problems in supernova theory. \acknowledgments This work has been partially supported by the MEC grants AYA2007-66256 and AYA2008-04211-C02-02, by the European Union FEDER funds, by the Generalitat de Catalunya, and by the ASI-INAF I/016/07/0. CB thanks Benoziyo Center for Astrophysics for support \bibliographystyle{aa}
\section{Introduction } Much attention has long been devoted to finding macroscopic magnetic ordering phenomena in organic materials. Experimentally, ferromagnetism has been discovered in pure carbon systems, such as carbon foam, graphite, oxidized C$_{60}$ and polymerized rhombohedral C$_{60}$ \cite{r1,r2,r3,r4,r5}, which has stimulated renewed interests in their fundamental importance and potential applications in high-technology, e.g., the spintronics. Much theoretical work has been done to study magnetism in nanographites \cite{r6,r7,r8,r9,r10}, C$_{60}$ polymers \cite{r11}, and all-carbon nanostructures \cite{r12,r13}. However, the microscopic origin of ferromagnetism remains controversial. The recent proton irradiation experiments in graphite\cite{r2,r3} have shown the importance of hydrogen in inducing the magnetization instead of magnetic impurities. Disorder induced by He$^{+}$ ions irradiation does not produce such a large magnetic moment as obtained with protons \cite{r14}. Theoretically, the possible magnetism arising from the adsorption of hydrogen atom on graphite has been studied \cite{r9,r10}. It can be easily speculated that hydrogen can trigger the $sp^2$-$sp^3$ transformation, promoting the magnetic ordering in other carbon structures, especially carbon nanotubes with a surface of positive curvature. Herein, we focus on the electronic and magnetic properties of single-walled carbon nanotubes (SWNTs) with hydrogen atoms adsorbed on their surfaces. Our results show that hydrogenated carbon nanotubes are on the verge of magnetism instability, and the \emph{combination} of charge transfer and carbon network distortion drives flat-band ferromagnetism. To our knowledge, this is the first comprehensive $ab$ $initio$ study on the physical origin of flat-band ferromagnetism in the real carbon nanotube materials. Moreover, the applied strain and external electric field are found to have a strong influence on the flat-band spin-splitting, resulting in the variation of spin-relevant physical properties. \section{Results and discussion} \begin{figure}[htbp] \includegraphics[width=0.5\columnwidth, angle=-90]{figure1.eps} \label{fig1} \caption{Schematic geometrical structures of hydrogenated zigzag (\emph{top panels}) and armchair (\emph{bottom panels}) tubes in the higher hydrogen concentration A with one hydrogen per tube period, or the lower hydrogen concentration B with one hydrogen per every two tube periods. Blue (grey) balls represent the hydrogen (carbon) atoms.} \end{figure} \begin{figure*}[htbp] \includegraphics[width=1.9\columnwidth]{figure2.eps} \label{fig2} \caption{Spin-polarized LDA (\emph{left panels}) and GGA (\emph{right panels}) ground state band structures of hydrogenated zigzag (8,0) and (9,0), and armchair (5,5) and (10,10) tubes in the two different systems A (\emph{top panels}) and B (\emph{bottom panels}). Insets show the band structures around $\epsilon _{F}$ in an enlarged energy scale. The Fermi level is set at zero. Spin-up and spin-down channels are represented by black and cyan, respectively.} \end{figure*} Instead of the simple $sp^2$ bonds in graphite, the bonds in carbon nanotubes are of $sp^2$-$sp^3$ character due to the tube's curvature effect, making the hybridization of $\sigma $, $\sigma ^\ast $, $\pi $ and $\pi ^\ast $ orbitals quite larger, especially for small-diameter tubes. Thus the magnetic property of hydrogenated SWNTs are more complicated than that of hydrogenated graphite, with a large dependence on the radii, the chiralities, and hydrogen concentration. Here, two types of linear hydrogen concentration A and B are shown in Fig. 1. There is one hydrogen atom per tube period in the higher concentration A, and one hydrogen atom per every two tube periods in the lower concentration B. The higher H concentration leads to a larger structure deformation, as can be seen in Fig. 1. Stable C-H bond length is 1.1 {\AA}, typical of covalent bonding (cf. 1.09 {\AA} in methane). Recently, the cooperative alignment of the absorbed atoms has been observed in graphene experimentally \cite{r15}. In the case of carbon nanotubes, the absorption of hydrogen on tube wall is easier than that on graphene due to the curvature effect \cite{r16}. Moreover, the adatoms' cooperative alignment can be enhanced by high curvature regions of nanotube \cite{r16}, that can result from pressure \cite{r17}, compression transverse to its axis \cite{r16,r18}, or tube-substrate interaction \cite{r19}. Now, let us study electronic structures of hydrogenated SWNTs, taking the zigzag (8,0), (9,0), armchair (5,5) and (10,10) tubes as examples. Fig. 2 presents their ground state band structures obtained by using spin-polarized LDA (\emph{left panels}) and GGA (PBE exchange correlation functional) (\emph{right panels}). In spin-unpolarized LDA and GGA paramagnetic (PM) band structures (not shown here completely), a common feature is that hydrogen atom induces a half-filled flat-band in the tube's energy gap around $\epsilon _{F}$ due to the odd electrons in compounds, as seen in the LDA-PM ground state band structure of (5,5)-B in Fig. 2. Apparently, flat-band causes an extremely high density of states around $\epsilon _{F}$, and if the Coulomb interaction between itinerant electrons in the band is introduced, magnetic instability would occur. It has been shown that flat-band leads to ferromagnetism for certain models \cite{r20}. In hydrogenated SWNTs, the spin-spin interaction plays a similar role as the Coulomb one, and lifts the spin degeneracy of the flat band. As a result, the flat-band's spin-splitting magnitude and the energy position relative to $\epsilon _{F}$ determine the ground state properties of system. As we can see in the right panels of Fig. 2, the spin-splitting magnitude has increased after introducing the generalized gradient correction in the exchange correlation functional, compared to the LDA without correction (\emph{left panels}), which makes ferromagnetic (FM) state become more favorable in energy. The adsorption of hydrogen atom is found to hardly affect the gaps of zigzag (8,0) and (9,0) tubes \cite{r21}, while a large tube's energy gap is opened in the metallic armchair (5,5) and (10,10) tubes. In the LDA results, (8,0)-A,B, (5,5)-A and (10,10)-A exhibit FM semiconducting characteristic, whereas (9,0)-A,B and (10,10)-B are FM metals. (5,5)-B has a PM metallic ground state under LDA. However, the enhanced spin-splitting in GGA makes (9,0)-A,B and (10,10)-B present a FM semiconducting behaviour, not metallic one under LDA. Furthermore, a FM metallic state is produced for (5,5)-B, not the PM metallic one obtained by LDA.\\ \begin{figure}[htbp] \includegraphics[width=0.8\columnwidth]{figure3.eps} \label{fig3} \caption{The 0.003 {\AA}$^{-3}$ magnetization density isosurfaces for the full-occupied spin-up GGA flat-band in (9,0)-B: (a) top view, and (b) front view. Spin-polarized LDA and GGA ground state band structures for (9,0)-B: (c) with only the location of hydrogen atom relaxed, (d) under the $4\%$ axial stretch strain, (e) under the $4\%$ axial compression strain.} \end{figure} In order to investigate the physical origin of flat-band ferromagnetism, we plot the spin-up density of the full-occupied GGA flat-band of (9,0)-B (black band, see Fig. 2) in Fig. 3. A $sp^3$-like hybridization is induced on the carbon atom attached to the hydrogen atom, leading to whole carbon network distortion, accompanied by charge transfer from hydrogen to the bonded carbon atom (0.352 and 0.334 electron for LDA and GGA (PBE), respectively). Magnetism is strong itinerant in both circumferential and tube axial directions, arising from the hybridization of H-$s$ orbital with tube-$\pi$ orbitals. Magnetic moment per unit cell is 0.63 and 0.96 $\mu _B $ in LDA and GGA, respectively. Obviously, this $s$-$\pi$ hybridization is relevant to structure distortion and affects the flat-band spin-splitting magnitude. To get such an insight, we carried out both LDA and GGA calculations on the undistorted (9,0)-B structure with only the location of hydrogen atom relaxed. Charge transfer still exists, and the band structures plotted in Fig. 3(c) clearly show a disappearance of the flat-band spin-splitting and also magnetism. Weak magnetic moment 0.2 $\mu_B $ per unit cell appears even if on-site Coulomb repulsion U = 3.0 eV \cite{r22} is considered in the LDA+U calculation. These results indicate that only charge transfer is not sufficient to induce a large magnetic moment, therefore structure distortion is indispensable. As a result, we can anticipate that the applied strain could tune the physical properties of hydrogenated SWNTs by affecting the spin-splitting of flat-band. We applied a $4\%$ stretch or compression strain along the tube axis, then the atomic positions were optimized again. Corresponding ground state electronic structures are given in Figs. 3(d)-3(e). As expected, strain do has an important effect on the magnetic ground state properties. The $4\%$ axial stretch strain causes the length of C-C bonds, beneath the H atom, further stretched and deviate from the regular value of 1.42 \AA, which enlarges the energy gap, as shown in Fig. 3(d). Even a metal-insulator transition is found in LDA band structure, accompanied by an increase in the magnetic moment to 0.96 $\mu _B$ (normally 0.63 $\mu _B$). Under $4\%$ axial compression strain, the compound presents the metallic or semiconducting characteristic similar to those without strain in both LDA and GGA, but the magnetic moment in LDA is reduced to 0.53 $\mu _B $. Under both axial strains, the transferred charge remains to be 0.35 $\sim $ 0.36 electron in LDA and 0.33 $\sim $ 0.34 electron in GGA, proving that carbon network distortion plays a very important role in the flat-band spin-splitting. Of course, flat-band and magnetism would also disappear if we remove the hydrogen atom from the distorted SWNTs, indicating the necessity of hydrogen in inducing the ferromagnetic ordering, which is similar to the role of so called ``carbon radicals" in magnetic all-carbon structures \cite{r7}. Fig. 4 summaries structure information and ground state properties of hydrogenated zigzag ($n$,0) ($n$=6--12) and armchair ($n$,$n$) ($n$=5--11) tubes in both A and B cases (see Fig. 1). Generally speaking, magnetic moment per unit cell has increased gradually with the change of tube index $n$ (\emph{top panels}), and approaches a saturation value in both LDA and GGA results. In ($n$,0)-A case, magnetic moment curve presents oscillation behavior (especially in LDA) that might be relevant to periodic change of the band gap in zigzag-type tubes. The higher concentration A-type compounds have a larger magnetic moment than that of the lower concentration B-type compounds, especially in smaller diameter tubes. In the bottom panels of Fig. 4, the curves of energy difference ($E_{PM}-E_{FM}$) between FM and PM states per unit cell vs tube index $n$ present a similar trend as found in the curves of magnetic moment vs $n$. Introducing of the generalized gradient correction enhances the spin-splitting, as discussed for Fig. 2, which leads that the FM state becomes more favorable in GGA than in LDA, evidenced by the energy difference $E_{PM}-E_{FM}$. These reveal that magnetic properties of hydrogenated carbon nanotubes are affected by the structure characteristic of the SWNTs host, and the concentration of hydrogen. \begin{figure}[htbp] \includegraphics[width=0.9\columnwidth]{figure4.eps} \label{fig4} \caption{From the top down, the evolution of magnetic moment per unit cell and energy difference ($E_{PM}-E_{FM}$) per unit cell vs the tube index $n$. Left panels are from spin-polarized LDA, and right panels are from spin-polarized GGA (PBE).} \end{figure} Finally, we also investigate the effect of transverse electric fields in the Y direction of cross section (see Fig. 3(a)). Absorption of H atom makes SWNT become a polar organic compound. Here, the polar C-H bond is just along the Y direction. It can be anticipated that charge redistribution driven by an external electric field could occur, which would induce a subtle change of electronic structure. Both LDA and GGA simulations on (10,10)-A reveal the spin-dependent effect of applied external electric fields with a magnitude of 0.1 $\sim$ 0.4 V/{\AA}, illustrated in Fig. 5. Compared with Fig. 2, the applied field gradually decreases the gap between spin-up and -down flat band at 0.1 and 0.2 V/{\AA}, finally closes it and directly drives an insulator-metal transition at 0.3 V/{\AA} in LDA and 0.4 V/{\AA} in GGA. This effect is different from the case of graphene nanoribbons \cite{r6}, where the band gap closure takes place only for one spin channel. The fantastic ``beating" behaviour of spin-polarized flat-band under the external electric field can be used for designing quantum switches. \begin{figure}[htbp] \includegraphics[width=0.9\columnwidth]{figure5.eps} \label{fig5} \caption{Spin-polarized LDA (\emph{top panels}) and GGA (\emph{bottom panels}) ground state band structures of hydrogenated armchair (10,10)-A with the applied electric fields in the modulus of 0.1$\sim$0.4 V/{\AA} along the Y direction of cross section (see Fig. 3(a)). The Fermi Levels are marked in black dash lines.} \end{figure} \section{Conclusions} We have demonstrated how the \emph{combined effect} of charger transfer and carbon network distortion makes spin-polarized flat-band appear in the tube's energy gap, based on spin-polarized LDA and GGA (PBE) calculations, which is the origin of a variety of spin-relevant physical properties. Furthermore, hydrogenated SWNTs' ground state properties are found to depend largely on the radii, the chiralities of the SWNTs and the concentration of hydrogen. Applied strain and transverse electric field can effectively tune the flat-band spin-splitting strength, and even induce insulator-metal transition. Our results indicate that hydrogenated SWNTs are the promising candidates for the carbon-based nanometer-ferromagnetism, spintronic devices, and even quantum switches. Also, the intrinsic ferromagnetism mechanism revealed in our paper can be applied to hydrogenated graphene and other carbon-based material with extended surface.\\ \section{Theoretical methods and Models} We carried out the numerical calculations using the Vienna $ab$ $initio$ Simulation Package (VASP) \cite{r23} within the frameworks of spin-polarized local density approximation (LDA) and generalized gradient approximation (GGA) (PBE exchange correlation functional) \cite{r24}. The ion-electron interaction was modeled by the projector augmented wave (PAW) method \cite{r25} with a uniform energy cutoff of 400 eV. We used periodic boundary conditions and a supercell large enough to prohibit the electronic and the dipole-dipole interactions between neighboring tubes. The spacing between $k$ points was 0.03 {\AA}$^{-1}$: 1$\times$1$\times$8 k-point sampling in ($n$,0)-A, 1$\times$1$\times$4 in ($n$,0)-B, 1$\times$1$\times$14 in ($n$,$n$)-A, and 1$\times$1$\times$7 in ($n$,$n$)-B. The geometrical structures of hydrogenated SWNTs were optimized by employing the conjugate gradient technique, and in the final geometry no force on the atoms exceeded 0.01 eV/{\AA}.
\section{Introduction} We consider an $n$-element set $X$ with an unknown total ordering $\leq$. The ordering can be accessed via an oracle that, given two elements $x,y\in X$, tells us whether $x<y$ or $x>y$. It is easily seen that the minimum element of $X$ can be found using $n-1$ comparisons. This is optimal in the sense that $n-2$ comparisons are not enough to find the minimum element in the worst case. One of the nice little surprises in computer science is that if we want to find \emph{both} the minimum and the maximum, we can do significantly \emph{better} than finding the minimum and the maximum separately. Pohl~\cite{Pohl72} proved that $\lceil 3n/2 \rceil -2$ is the optimal number of comparisons for this problem ($n\geq 2$). The algorithm first partitions the elements of $X$ into pairs and makes a comparison in each pair. The minimum can then be found among the ``losers'' of these comparisons, while the maximum is found among the ``winners.'' Here we consider the problem of determining both the minimum and the maximum in the case where the oracle is not completely reliable: it may sometimes give a false answer, but only at most $k$ times during the whole computation, where $k$ is a given parameter. We refer to this model as \emph{computation against $k$ lies}. Let us stress that we admit repeating the same query to the oracle several times, and each false answer counts as a lie. This seems to be the most sensible definition---if repeated queries were not allowed, or if the oracle could always give the wrong answer to a particular query, then the minimum cannot be determined. So, for example, if we repeat a given query $2k+1$ times, we always get the correct answer by majority vote. Thus, we can simulate any algorithm with a reliable oracle, asking every question $2k+1$ times, but for the problems considered here, this is not a very efficient way, as we will see. The problem of finding both the minimum and the maximum against $k$ lies was investigated by Aigner~\cite{Aigner97}, who proved that $(k+\mathrm{O}(\sqrt{k}))n$ comparisons always suffice.\footnote{Here and in the sequel, $\mathrm{O}(.)$ and $\mathrm{\Omega}(.)$ hide only absolute constants, independent of both $n$ and~$k$.} We improve on this as follows. \begin{theorem}\label{thm:1} There is an algorithm that finds both the minimum and the maximum among $n$ elements against $k$ lies using at most $(k+1+C)n+\mathrm{O}(k^3)$ comparisons, where $C$ is a constant. \end{theorem} Our proof yields the constant $C$ reasonably small (below $10$, say, at least if $k$ is assumed to be sufficiently large), but we do not try to optimize it. \heading{Lower bounds. } The best known lower bounds for the number of comparisons necessary to determine both the minimum and the maximum against $k$ lies have the form $(k+1+c_k)n-D$, where $D$ is a small constant and the $c_k$ are as follows: \begin{itemize} \item $c_0=0.5$, and this is the best possible. This is the result of Pohl~\cite{Pohl72} for a truthful oracle mentioned above. \item $c_1=\frac{23}{32}=0.71875$, and this is again tight. This follows from a recent work by Gerbner, P\'alv\"olgyi, Patk\'os, and Wiener~\cite{GPPW} who determined the optimum number of comparisons for $k=1$ up to a small additive constant: it lies between $\lceil \frac{87}{32}n \rceil -3$ and $\lceil \frac{87}{32}n \rceil +12$. This proves a conjecture of Aigner~\cite{Aigner97}. \item $c_k=\mathrm{\Omega}(2^{-5k/4})$ for all $k$, as was shown by Aigner~\cite{Aigner97}. \end{itemize} The optimal constant $c_1=\frac{23}{32}$ indicates that obtaining precise answers for $k>1$ may be difficult. \heading{Related work. } The problem of determining the minimum alone against $k$ lies was resolved by Ravikumar, Ganesan, and Lakshmanan~\cite{RGL87}, who proved that finding the minimum against $k$ lies can be performed by using at most $(k+1)n-1$ comparisons, and this is optimal in the worst case. The problem considered in this paper belongs to the area of \emph{searching problems against lies} and, in a wider context, it is an example of ``computation in the presence of errors.'' This field has a rich history and beautiful results. A prototype problem, still far from completely solved, is the \emph{R\'enyi--Ulam liar game} from the 1960s, where one wants to determine an unknown integer $x$ between $1$ and $n$, an oracle provides comparisons of $x$ with specified numbers, and it may give at most $k$ false answers. We refer to the surveys by Pelc~\cite{Pelc} and by Deppe~\cite{Deppe} for more information. \section{A simple algorithm} Before proving Theorem~\ref{thm:1}, we explain a simpler algorithm, which illustrates the main ideas but yields a weaker bound. We begin with formulating a generic algorithm, with some steps left unspecified. Both the simple algorithm in this section and an improved algorithm in the next sections are instances of the generic algorithm. \framedpar{ \begin{center}\bf The generic algorithm \end{center} \begin{enumerate} \item \label{alg1:partition} For a suitable integer parameter $s=s(k)$, we arbitrarily partition the considered $n$-element set $X$ into $n/s$ groups $X_1,\ldots,X_{n/s}$ of size $s$ each.\footnote{If $n$ is not divisible by $s$, we can form an extra group smaller than $s$ and treat it separately, say---we will not bore the reader with the details.} \item \label{alg1:sort} In each group $X_i$, we find the minimum $m_i$ and the maximum $M_i$. The method for doing this is left unspecified in the generic algorithm. \item \label{alg1:find} We find the minimum of $\{m_1,\ldots,m_{n/s}\}$ against $k$ lies, and independently, we find the maximum of $\{M_1,M_2,\ldots,M_{n/s}\}$ against $k$ lies. \end{enumerate} } The correctness of the generic algorithm is clear, provided that Step~\ref{alg1:sort} is implemented correctly. Eventually, we set $s:=k$ in the simple and in the improved algorithm. However, we keep $s$ as a separate parameter, because the choice $s:=k$ is in a sense accidental. In the simple algorithm we implement Step~\ref{alg1:sort} as follows. \medskip \framedpar{ \begin{center}\bf Step \ref{alg1:sort} in the simple algorithm \end{center} \begin{enumerate} \renewcommand{\theenumi}{\ref{alg1:sort}.\arabic{enumi}} \item \label{again} (Sorting.) We sort the elements of $X_i$ by an asymptotically optimal sorting algorithm, say mergesort, using $\mathrm{O}(s\log s)$ comparisons, and ignoring the possibility of lies. Thus, we obtain an ordering $x_1,x_2,\ldots,x_s$ of the elements of $X_i$ such that \emph{if} all queries during the sorting have been answered correctly, \emph{then} $x_1< x_2< \cdots < x_s$. If there was at least one false answer, we make no assumptions, except that the sorting algorithm does not crash and outputs some ordering. \item \label{verify} (Verifying the minimum and maximum.) For each $j=2,3,\ldots,s$, we query the oracle $k+1$ times with the pair $x_{j-1},x_{j}$. If any of these queries returns the answer $x_{j-1}>x_{j}$, we \emph{restart}: We go back to Step~\ref{again} and repeat the computation for the group $X_i$ from scratch. Otherwise, if all the answers are $x_{j-1}<x_{j}$, we proceed with the next step. \item\label{allright} We set $m_i:=x_1$ and $M_i:=x_s$. \end{enumerate} } \medskip \begin{lemma}[Correctness] The simple algorithm always correctly computes the minimum and the maximum against $k$ lies. \end{lemma} \begin{proof} We note that once the processing of the group $X_i$ in the above algorithm reaches Step~\ref{allright}, then $m_i=x_1$ has to be the minimum. Indeed, for every other element $x_j$, $j\ge 2$, the oracle has answered $k+1$ times that $x_j>x_{j-1}$, and hence $x_j$ cannot be the minimum. Similarly, $M_i$ has to be the maximum, and thus the algorithm is always correct. \end{proof} Actually, at Step~\ref{allright} we can be sure that $x_1,\ldots,x_s$ is the sorted order of $X_i$, but in the improved algorithm in the next section the situation will be more subtle. The next lemma shows, that the simple algorithm already provides an improvement of Aigner's bound of $(k+\mathrm{O}(\sqrt k))n$. \begin{lemma}[Complexity] The number of comparisons of the simple algorithm for $s=k$ on an $n$-element set is $(k+\mathrm{O}(\log k))n + \mathrm{O}(k^3)$. \end{lemma} \begin{proof} For processing the group $X_i$ in Step~\ref{alg1:sort}, we need $\mathrm{O}(s\log s)+(k+1)(s-1)=k^2+\mathrm{O}(k\log k)$ comparisons, provided that no restart is required. But since restarts may occur only if the the oracle lies at least once, and the total number of lies is at most $k$, there are no more than $k$ restarts for all groups together. These restarts may account for at most $k(k^2+\mathrm{O}(k\log k))= \mathrm{O}(k^3)$ comparisons. Thus, the total number of comparisons in Step~\ref{alg1:sort} is $\frac ns(k^2+\mathrm{O}(k\log k))+\mathrm{O}(k^3)= (k+\mathrm{O}(\log k))n+\mathrm{O}(k^3)$. As we mentioned in the introduction, the minimum (or maximum) of an $n$-element set against $k$ lies can be found using $(k+1)n-1$ comparisons, and so Step~\ref{alg1:find} needs no more than $2(k+1)(n/s)=\mathrm{O}(n)$ comparisons. (We do not really need the optimal algorithm for finding the minimum; any $\mathrm{O}((k+1)n)$ algorithm would do.) The claimed bound on the total number of comparisons follows. \end{proof} \section{The improved algorithm: Proof of Theorem~\ref{thm:1}} In order to certify that $x_1$ is indeed the minimum of $X_i$, we want that for every $x_j$, $j\ne 1$, the oracle declares $x_j$ larger than some other element $k+1$ times. (In the simple algorithm, these $k+1$ comparisons were all made with $x_{j-1}$, but any other smaller elements will do.) This in itself requires $(k+1)(s-1)$ queries per group, or $(k+1)(n-n/s)$ in total, which is already close to our target upper bound in Theorem~\ref{thm:1} (we note that $s$ has to be at least of order $k$, for otherwise, Step~\ref{alg1:find} of the generic algorithm would be too costly). Similarly, every $x_j$, $j\ne s$, should be compared with smaller elements $k+1$ times, which again needs $(k+1)(n-n/s)$ comparisons, so all but $\mathrm{O}(n)$ comparisons in the whole algorithm should better be used for \emph{both} of these purposes. In the simple algorithm, the comparisons used for sorting the groups in Step~\ref{again} are, in this sense, wasted. The remedy is to use most of them also for verifying the minimum and maximum in Step~\ref{verify}. For example, if the sorting algorithm has already made comparisons of $x_{17}$ with 23 larger elements, in the verification step it suffices to compare $x_{17}$ with $k+1-23$ larger elements. One immediate problem with this appears if the sorting algorithm compares $x_{17}$ with some $b>k+1$ larger elements, the extra $b-(k+1)$ comparisons are wasted. However, for us, this will not be an issue, because we will have $s=k$, and thus each element can be compared to at most $k-1$ others (assuming, as we may, that the sorting algorithm does not repeat any comparison). Another problem is somewhat more subtle. In order to explain it, let us represent the comparisons made in the sorting algorithm by edges of an ordered graph. The vertices are $1,2,\ldots,s$, representing the elements $x_1,\ldots,x_s$ of $X_i$ in sorted order, and the edges correspond to the comparisons made during the sorting, see the figure below on the left. \iepsfig{aftersort} In the verification step, we need to make additional comparisons so that every $x_j$, $j\ne 1$, has at least $k+1$ comparisons with smaller elements and every $x_j$, $j\ne s$, has at least $k+1$ comparisons with larger elements. This corresponds to adding suitable extra edges in the graph, as in the right drawing above (where $k=2$, and the added edges are drawn on the bottom side). As the picture illustrates, sometimes we cannot avoid comparing some element with \emph{more} than $k+1$ larger ones or $k+1$ smaller ones (and thus some of the comparisons will be ``half-wasted''). For example, no matter how we add the extra edges, the elements $x_1,x_2,x_3$ together must participate in at least $3$ half-wasted comparisons. Indeed, $x_2$ and $x_3$ together require $6$ comparisons to the left (i.e.\ with a smaller element). These comparisons can be ``provided'' only by $x_1$ and $x_2$, which together want only 6 comparisons to the right---but $3$ of these comparisons to the right were already made with elements larger than $x_3$ (these are the arcs intersecting the dotted vertical line in the picture). The next lemma shows that this kind of argument is the only source of wasted comparisons. For an ordered multigraph $H$ on the vertex set $\{1,2,\ldots,s\}$ as above, let us define $t(H)$, the \emph{thickness} of $H$, as $\max\{t(j): j=2,3,\ldots,s-1\}$, where $t(j):=|\{\{a,b\}\in E(H): a<j<b\}|$ is the number of edges going ``over'' the vertex~$j$. \begin{lemma}\label{l:addedges} Let $H$ be an undirected multigraph without loops on $\{1,2,\ldots,s\}$ such that for every vertex $j=1,2,\ldots,s$, \begin{align*} d_H^{{\rm left}}(j) & := |\{ \{i,j\}\in E(H) : i<j\}| \leq k+1\,,\\ d_H^{{\rm right}}(j) & := |\{ \{i,j\}\in E(H) : i>j\}| \leq k+1\,. \end{align*} Then $H$ can be extended to a multigraph $\overline H$ by adding edges, so that \begin{enumerate} \item[\rm(i)] every vertex $j\ne 1$ has at least $k+1$ left neighbors and every vertex $j\ne s$ has at least $k+1$ right neighbors; and \item[\rm(ii)] the total number of edges in $\overline H$ is at most $(k+1)(s-1) + t(H)$. \end{enumerate} \end{lemma} The proof is a network flow argument and therefore constructive. We postpone it to the end of this section. For a comparison-based sorting algorithm $\mathcal{A}$, we define the \emph{thickness} $t_\mathcal{A}(s)$ in the natural way: It is the maximum, over all $s$-element input sequences, of the thickness $t(H)$ of the corresponding ordered graph $H$ (the vertices of $H$ are ordered as in the output of the algorithm and each comparison contributes to an edge between its corresponding vertices). As the above lemma shows, the number of comparisons used for the sorting but not for the verification can be bounded by the thickness of the sorting algorithm. \begin{lemma}\label{l:quicksort} There exists a (deterministic) sorting algorithm $\mathcal{A}$ with thickness $t_{\mathcal{A}}(s)=\mathrm{O}(s)$. \end{lemma} \begin{proof} The algorithm is based on Quicksort, but in order to control the thickness, we want to partition the elements into two groups of equal size in each recursive step. We thus begin with computing the median of the given elements. This can be done using $\mathrm{O}(s)$ comparisons (see, e.g., Knuth~\cite{Knuth1973}; the current best deterministic algorithm due to Dor and Zwick~\cite{DZ99} uses no more than $2.95s +\mathrm{o}(s)$ comparisons). These algorithms also divide the remaining elements into two groups, those smaller than the median and those larger than the median. To obtain a sorting algorithm, we simply recurse on each of these groups. The thickness of this algorithm obeys the recursion $t_{\mathcal{A}}(s)\le \mathrm{O}(s)+t_{\mathcal{A}}(\lfloor s/2\rfloor))$, and thus it is bounded by $\mathrm{O}(s)$. \end{proof} We are going to use the algorithm $\mathcal{A}$ from the lemma in the setting where some of the answers of the oracle may be wrong. Then the median selection algorithm is not guaranteed to partition the current set into two groups of the same size and it is not sure that the running time does not change. However, we can check if the groups have the right size and if the running time does not increase too much. If some test goes wrong, we restart the computation (similar to the simple algorithm). Now we can describe the improved algorithm, again by specifying Step~\ref{alg1:sort} of the generic algorithm. \medskip \framedpar{ \begin{center}\bf Step~\ref{alg1:sort} in the improved algorithm \end{center} \begin{enumerate} \renewcommand{\theenumi}{\ref{alg1:sort}.\arabic{enumi}$'$} \item \label{again2} (Sorting.) We sort the elements of $X_i$ by the algorithm $\mathcal{A}$ with thickness $\mathrm{O}(s)$ as in Lemma~\ref{l:quicksort}. If an inconsistency is detected (as discussed above), we restart the computation for the group $X_i$ from scratch. \item \label{verify2} (Verifying the minimum and maximum.) We create the ordered graph $H$ corresponding to the comparisons made by $\mathcal{A}$, and we extend it to a multigraph $\overline{H}$ according to Lemma~\ref{l:addedges}. We perform the comparisons corresponding to the added edges. If we encounter an inconsistency, then we restart: We go back to Step~\ref{again2} and repeat the computation for the group $X_i$ from scratch. Otherwise, we proceed with the next step. \item\label{allright2} We set $m_i:=x_1$ and $M_i:=x_s$. \end{enumerate} } \medskip \begin{proof}[Proof of Theorem~\ref{thm:1}] The correctness of the improved algorithm follows in the same way as for the simple algorithm. In Step \ref{verify2}, the oracle has declared every element $x_j$, $j\ne 1$, larger than some other element $k+1$ times, and so $x_j$ cannot be the minimum. A similar argument applies for the maximum. It remains to bound the number of comparisons. From the discussion above, the number of comparisons is at most $((k+1)(s-1)+t_{\mathcal{A}}(s))(\frac{n}{s}+k)+2(k+1)\frac{n}{s}$, with $t_{\mathcal{A}}(s)=\mathrm{O}(s)$. For $s=k$, we thus get that the number of comparisons at most $(k+1+C)n+\mathrm{O}(k^3)$ for some constant $C$, as claimed. \end{proof} \begin{proof}[Proof of Lemma~\ref{l:addedges}] We will proceed in two steps. First, we construct a multigraph $H^*$ from $H$ by adding a maximum number of (multi)edges such that the left and right degree of every vertex are still bounded above by $k+1$. Second, we extend $H^*$ to $\overline{H}$ by adding an appropriate number of edges to each vertex so that condition (i) holds. For an ordered multigraph $H'$ on $\{1,2,\ldots,s\}$ with left and right degrees upper bounded by $k+1$, let us define the \emph{defect} $\Delta(H')$ as \[ \Delta(H') := \sum_{j=1}^{s-1} (k+1-d_{H'}^{\mathrm{right}}(j))+ \sum_{j=2}^{s} (k+1-d_{H'}^{\mathrm{left}}(j))\,. \] We have $\Delta(H')=2(k+1)(s-1)-2e(H')$, where $e(H')$ is the number of edges of~$H'$. By a network flow argument, we will show that by adding suitable $m^*:= (k+1)(s-1)-e(H)-t(H)$ edges to $H$, one can obtain a multigraph $H^*$ in which all left and right degrees are still bounded by $k+1$ and such that $\Delta(H^*) = 2t(H)$. The desired graph $\overline H$ as in the lemma will then be obtained by adding $\Delta(H^*)$ more edges: For example, for every vertex $j\ge 2$ of $H^*$ with $d_{H^*}^{\rm left}(j)<k+1$, we add $k+1-d_{H^*}^{\rm left}(j)$ edges connecting $j$ to~$1$, and similarly we fix the deficient right degrees by adding edges going to the vertex~$s$. It remains to construct $H^*$ as above. To this end, we define an auxiliary directed graph $G$, where each directed edge $e$ is also assigned an integral capacity $c(e)$; see \figurename~\ref{fig:flow1}. \begin{figure}[htbp] \centering \subfigure[The graph $G$ with capacities.]{ \label{fig:flow1} \includegraphics[scale=.95]{flow1}\hspace{0.5cm}} \subfigure[The cut $S_2$ in $G$.]{ \label{fig:flow2} \includegraphics[scale=.95]{flow2}} \caption{The directed graph $G$ constructed in the proof of Lemma~\ref{l:addedges}.} \end{figure} The vertex set of $G$ consists of a vertex $j^-$ for every $j\in \{1,2,\ldots,s\}$, a vertex $j^+$ for every $j\in \{1,2,\ldots,s\}$, and two special vertices $a$ and $b$. There is a directed edge in $G$ from $a$ to every vertex $j^+$ and the capacity of this edge is $k+1-d_H^{\mathrm{right}}(j)$. Similarly, there is a directed edge in $G$ from every vertex $j^-$ to $b$, and the capacity of this edge is $k+1-d_H^{\mathrm{left}}(j)$. Moreover, for every $i,j$ with $1\le i<j\le s$, we put the directed edge $(i^+,j^-)$ in $G$, and the capacity of this edge is $\infty$ (i.e., a sufficiently large number). We will check that there is an integral $a$-$b$ flow in $G$ with value $m^*$ in $G$. By the max-flow min-cut theorem~\cite{FF56}, it suffices to show that every $a$-$b$ cut in $G$ has capacity at least~$m^*$ and there is an $a$-$b$ cut in $G$ with capacity $m^*$. Let $S\subseteq V(G)$ be a minimum $a$-$b$ cut. Let $i$ be the smallest integer such that $i^+\in S$. Since the minimum cut cannot use an edge of unbounded capacity, we have $j^- \in S$ for all $j>i$. We may assume without loss of generality that $j^+ \in S$ for all $j>i$ and $j^-\not\in S$ for all $j\leq i$ (the capacity of the cut does not decrease by doing otherwise). Therefore it suffices to consider $a$-$b$ cuts of the form \[ S_i := \{a\} \cup \{x^+ : x\geq i \} \cup \{x^- : x > i\} \] for $i=1,\ldots ,s$. The capacity of $S_i$, see \figurename~\ref{fig:flow2}, equals \[ \sum_{j<i} c(a,j^+) + \sum_{j > i} c(j^-,b) = (s-1)(k+1) -\sum_{j<i} d_H^{\mathrm{right}}(j) - \sum_{j>i} d_H^{\mathrm{left}}(j)\,. \] Now let us look at the quantity $\sum_{j<i} d^{\mathrm{right}}(j) +\sum_{j>i} d^{\mathrm{left}}(j)$, and see how much an edge $\{j,j'\}$ ($j<j'$) of $H$ contributes to it: For $j<i<j'$, the contribution is $2$, while all other edges contribute~$1$. Hence the capacity of the cut $S_i$ is $(k+1)(s-1)-e(H)-t(i)$, and the minimum capacity of an $a$-$b$-cut is $(k+1)(s-1)-e(H)-t(H)=m^*$ as required. Thus, there is an integral flow $f$ with value $m^*$ as announced above. We now select the edges to be added to $H$ as follows: For every directed edge $(i^+,j^-)$ of $G$, we add $f(i^+,j^-)$ copies of the edge $\{i,j\}$, which yields the multigraph $H^*$. The number of added edges is $m^*$, the value of the flow $f$, and the capacity constraints guarantee that all left and right degrees in $H^*$ are bounded by $k+1$. Moreover, the defect of $H^*$ is at most $2t(H)$. \end{proof} \section{Concluding remarks} We can cast the algorithm when $k=0$ sketched in the introduction into the framework of our generic algorithm. Namely, if we set $s=2$ and in Step~\ref{alg1:sort} we just compare the two elements in each group, then we obtain that algorithm. The main feature of our algorithm is that every restart only spoils one group. This allows us to keep the effect of lies local. In order to improve the upper bound of Theorem~\ref{thm:1} by the method of this paper, we would need a sorting algorithm with thickness $\mathrm{o}(s)$. (Moreover, to make use of the sublinear thickness, we would need to choose $s$ superlinear in $k$, and thus the sorting algorithm would be allowed to compare every element with only $\mathrm{o}(s)$ others.) The following proposition shows, however, that such a sorting algorithm does not exist. Thus, we need a different idea to improve Theorem \ref{thm:1}. \begin{proposition} Every (randomized) algorithm to sort an $s$-element set has thickness $\mathrm{\Omega}(s)$ in expectation. \end{proposition} \begin{proof} By Yao's principle \cite{Yao}, it is enough to show that every deterministic sorting algorithm $\mathcal{A}$ has expected thickness $\mathrm{\Omega}(s)$ for a random input. In our case, we assume that the unknown linear ordering of $X$ is chosen uniformly at random among all the $s!$ possibilities. In each step, the algorithm $\mathcal{A}$ compares some two elements $x,y\in X$. Let us say that an element $x\in X$ is \emph{virgin} at the beginning of some step if it hasn't been involved in any previous comparison, and elements that are not virgin are \emph{tainted}. A comparison is \emph{fresh} if it involves at least one virgin element. For notational convenience, we assume that $s$ is divisible by 8. Let $L\subset X$ consist of the first $s/2$ elements in the (random) input order (which is also the order of the output of the algorithm), and let $R:=X\setminus L$. Let $E_i$ be the event that the $i$th fresh comparison is an \emph{$LR$-comparison}, i.e., a comparison in which one of the two compared elements $x,y$ lies in $L$ and the other in $R$. We claim that for each $i=1,2,\ldots,s/8$, the probability of $E_i$ is at least $\frac13$. To this end, let us fix (arbitrarily) the outcomes of all comparisons made by $\mathcal{A}$ before the $i$th fresh comparison, which determines the set of tainted elements, and let us also fix the positions of the tainted elements in the input ordering. We now consider the probability of $E_i$ \emph{conditioned} on these choices. The key observation is that the virgin elements in the input ordering are still randomly distributed among the remaining positions (those not occupied by the tainted elements). Let $\ell$ be the number of virgin elements in $L$ and $r$ the number of virgin elements in $R$; we have $s/4\le \ell,r\le s/2$. We distinguish two cases. First, let only one of the elements $x,y$ compared in the $i$th fresh comparison be virgin. Say that $x$ is tainted and lies in $L$. Then the probability of $E_i$ equals $r/(\ell+r)\ge \frac13$. Second, let both of $x$ and $y$ be virgin. Then the probability of $E_i$ is $2\ell r/((\ell+r)(\ell+r-1))$, and since $s/4\le \ell,r\le s/2$, this probability exceeds~$\frac 49$. Thus, the probability of $E_i$ conditioned on \emph{every} choice of the outcomes of the initial comparisons and positions of the tainted elements is at least $\frac 13$, and so the probability of $E_i$ for a random input is at least $\frac 13$ as claimed. Thus, the expected number of $LR$-comparisons made by $\mathcal{A}$ is $\Omega(s)$. Let $a$ be the largest element of $L$, i.e., the $(s/2)$th element of $X$, and let $b$ be the smallest element of $R$, i.e., the $(s/2+1)$st element of $X$. Since we may assume that $\mathcal{A}$ doesn't repeat any comparison, there is at most one comparison of $a$ with $b$. Every other $LR$-comparison compares elements that have $a$ or $b$ (or both) between them. Thus, the expected thickness of $\mathcal{A}$ is at least half of the expected number of $LR$-comparisons, which is $\Omega(s)$. \end{proof} Note that the only thing which we needed in the proposition above was that the corresponding ordered graph is simple and has minimum degree at least $1$. \section*{Acknowledgments} We thank D\"om P\'alv\"olgyi for bringing the problem investigated in this paper to our attention. This work has been started at the 7th Gremo Workshop on Open Problems, Hof de Planis, Stels in Switzerland, July 6--10, 2009. We also thank the participants of the workshop for the inspiring atmosphere.
\section{introduction} \label{intro} Graphene\cite{geim,kim} | a single layer of carbon atoms arranged in a honeycomb lattice | has attracted huge attention in the physics community because of many unusual electronic, thermal and nanomechanical properties.\cite{reviews,beenakker:2008} In graphene the Fermi surface, at the charge neutrality point, reduces to two isolated points, the two inequivalent corners $K$ and $K'$ of the hexagonal Brillouin zone of the honeycomb lattice. In their vicinity the charge carriers form a gas of chiral massless quasiparticles with a characteristic conical spectrum. The low-energy dynamics is governed by the Dirac-Weyl (DW) equation\cite{semenoff,divincenzomele} in which the role of speed of light is played by the electron Fermi velocity. The chiral nature of the quasiparticles and their linear spectrum lead to remarkable consequences for a variety of electronic properties as weak localization, shot noise, Andreev reflection, and many others. Also the behavior in a perpendicular magnetic field discloses new physics. Graphene exhibits a zero-energy Landau level, whose existence gives rise to an unconventional half-integer quantum Hall effect, one of the peculiar hallmarks of the DW physics. Driven by the prospects of using this material in spintronic applications,\cite{spintronics,Trauzettel:2008} the study of spin transport is one of the most active field in graphene research.\cite{expspin1,expspin2,expspin3,expspin4,expspin5,expspin6} Several experiments have recently demonstrated spin injection, spin-valve effect, and spin-coherent transport in graphene, with spin relaxation length of the order of few micrometers.\cite{expspin2, expspin6} In this context a crucial role is played by the spin-orbit interaction. In graphene symmetries allow for two kinds of spin-orbit coupling (SOC).\cite{kane:2005} The {\em intrinsic} SOC originates from carbon intra-atomic SOC. It opens a gap in the energy spectrum and converts graphene into a topological insulator with a quantized spin-Hall effect.\cite{kane:2005} This term has been estimated to be rather weak in clean flat graphene.\cite{soguinea,so2,so3,so4} The {\em extrinsic} Rashba-like SOC originates instead from interactions with the substrate, presence of a perpendicular external electric field, or curvature of graphene membrane.\cite{soguinea,so2,so3,so5} This term is believed to be responsible for spin polarization\cite{rashbagraphene} and spin relaxation\cite{guinea2009,ertler:2009} physics in graphene. Optical-conductivity measurements could provide a way to determine the respective strength of both SOCs.\cite{philip} In this article we address the problem of ballistic spin-dependent scattering in the presence of inhomogeneous spin-orbit couplings. Our main motivation stems from a recent experiment that reported a large enhancement of Rashba SOC splitting in single-layer graphene grown on Ni(111) intercalated with a Au monolayer.\cite{varykhalov:2008} Further experimental results show that the intercalation of Au atoms between graphene and the Ni substrate is essential in order to observe sizable Rashba effect.\cite{rader:2009,varykhalov:2009} The preparation technique of Ref.~\onlinecite{varykhalov:2008} seems to provide a system with properties very close to those of freestanding graphene in spite of the fact that graphene is grown on a solid substrate. The presence of the substrate does not seem to fundamentally alter the electronic properties observed in suspended systems, \emph{i.e.}, the existence of Dirac points at the Fermi energy and the gapless conical dispersion in their vicinity. These results suggest that a certain degree of control on the SOC can be achieved by appropriate substrate engineering, with variations of the SOC strength on sub-micrometer scales, without spoiling the relativistic gapless nature of quasiparticles. This could pave the way for the realization of spin-optics devices for spin filtration and spin control for DW fermions in graphene. An optimal design would require a detailed understanding of the spin-resolved ballistic scattering through such \emph{spin-orbit nanostructures}, which is the aim of this paper. The problem of spin transport through nanostructures with inhomogeneous SOC has already been thoroughly studied in the case of two-dimensional electron gas in semiconductor heterostructures with Rashba SOC.\cite{marigliano1, khodas,marigliano2} Here the Rashba SOC~\cite{rashba:1960} | arising from the inversion asymmetry of the confinement potential | couples the electron momentum to the spin degree of freedom and thereby lifts the spin degeneracy. In this case, a region with finite SOC between two normal regions has properties similar to biaxial crystals: an electron wave incident from the normal region splits at the interface and the two resulting waves propagate in the SO region with different Fermi velocities and momenta.~\cite{marigliano1} This effect | analogous to the optical double-refraction | produces an interference pattern when the electron waves emerge in the second normal region. Moreover, electrons that are injected in an spin unpolarized state emerge from the SO region in a partially polarized state. Here we shall focus on the two simplest examples of SO nanostructures in graphene: (i) a single interface between two regions with different strengths of SOC; (ii) a SOC barrier, or double interface, \emph{i.e.}, a region of finite SOC in between two regions with vanishing SOC. Our analysis shows | in analogy to the case of 2DEG | that the ballistic propagation of carriers is governed by the spin-double refraction. We find that the scattering properties of the structure strongly depend on the injection angle. As a consequence, an initially unpolarized DW quasiparticle emerges from the SOC barrier with a finite spin polarization. In analogy to the edge states in the quantum spin-Hall effect,\cite{kane:2005} we also consider the possibility of edge states localized at the interface between regions with and without SOC. This paper is organized as follows. In Sec.~\ref{sec:2} we introduce the model and the transfer matrix formalism used in the rest of the paper. In Sec.~\ref{sec:3} we discuss the scattering problem at a single interface and the spectrum of edge states. In Sec.~\ref{NSON} we address the case of a double interface | a SOC barrier | and the final Sec.~\ref{sec:5} is devoted to the discussion of results and conclusions. \section{Model and Formalism} \label{sec:2} \begin{figure}[!t] \centering \includegraphics[width=\columnwidth]{figure1} \caption{\label{fig:kinetics} (Color online) Illustration of the kinematics of the scattering at a N-SO interface in graphene. The circles represent constant energy contours.} \end{figure} We consider a clean graphene sheet in the $xy$-plane with SOCs\cite{kane:2005,soguinea, rashbagraphene,stauber:2009,yamakage:2009} inhomogeneous along the $x$-direction. We shall restrict ourselves to a single-particle picture and neglect electron-electron interaction effects. The length scale over which the SOCs vary is assumed to be much larger than graphene's lattice constant ($a=0.246$~nm) but much smaller than the typical Fermi wavelength of quasiparticles $\lambda_\text{F}$. Since close to the Dirac points $\lambda_\text{F} \sim 1 / |E|$, at low energy $E$ this approximation is justified. This assumption ensures that we can use the continuum DW description, in which the two valleys are not coupled. Yet close to a Dirac point we can approximate the variation of SOCs as a sharp change. Focusing on a single valley, the single-particle Hamiltonian reads \begin{align}\label{eq:one} \mathcal{H} & = v_\text{F}\ {\bm \sigma} \cdot {\bf p} + \mathcal{H}_\text{SO},\\ \mathcal{H}_\text{SO} & = \frac{\lambda(x)}{2} \left( {\bm \sigma} \times {\bf s}\right)_z + \Delta(x) \sigma_z s_z,\label{eq:ham:so} \end{align} where $v_\text{F}\approx 10^6$ m/s is the Fermi velocity in graphene. In the following we set $\hbar =v_\text{F}=1$. The vector of Pauli matrices $\bm \sigma=(\sigma_x,\sigma_y)$ [resp. ${\bf s} =(s_x,s_y)$] acts in sublattice space [resp. spin space]. The term $\mathcal{H}_\text{SO}$ contains the extrinsic or Rashba SOC of strength $\lambda$ and the intrinsic SOC of strength $\Delta$. While experimentally the Rashba SOC can be enhanced by appropriate optimization of the substrate up to values of the order of $14$ meV,~\cite{varykhalov:2008} the intrinsic SOC seems at least two orders of magnitude smaller. Yet, the limit of large intrinsic SOC is of considerable interest since in this regime graphene becomes a topological insulator.\cite{kane:2005} Thus in this paper we shall not restrict ourself to the experimentally relevant regime $\lambda\gg \Delta$ but consider also the complementary regime. The wave function $\Psi$ is expressed as \begin{equation*} \Psi^\text{T}=(\Psi_{A\uparrow},\Psi_{B\uparrow}, \Psi_{A\downarrow},\Psi_{B\downarrow}), \end{equation*} where the superscript ${}^\text{T}$ denotes transposition. Spectrum and eigenspinors of the Hamiltonian (\ref{eq:one}) with uniform SOCs are briefly reviewed in Appendix \ref{app:so}. The spectrum consists of four branches $E_{\alpha,\epsilon}({\bf k})$ labelled by the two quantum numbers $\epsilon=\pm1$ and $\alpha=\pm1$. Here, the first distinguishes particle and hole branches, the second gives the sign of the expectation value of the spin projection along the in-plane direction perpendicular to the propagation direction ${\bf k}$. The spectrum strongly depends on the ratio \begin{equation}\label{eq:eta} \eta=\frac{\Delta}{\lambda}. \end{equation} For $\eta>1/2$ a gap separates particle and hole branches. The gap closes at $\eta=1/2$ and for $\eta<1/2$ one particle branch and one hole branch are degenerate at ${\bf k}=0$ (see Fig.~\ref{fig:spec} in App.~\ref{app:so}). We now briefly summarize the transfer matrix approach employed in this paper to solve the DW scattering problem.~\cite{mckellar:1987,peeters,luca1,luca2} We assume translational invariance in the $y$-direction, thus the scattering problem for the Hamiltonian (\ref{eq:one}) reduces to an effectively one-dimensional (1D) one. The wave function factorizes as $\Psi(x,y)=\text{e}^{\text{i} k_yy}\chi(x)$, where $k_y$ is the conserved $y$-component of the momentum, which parameterizes the eigenfunctions of the Hamiltonian of given energy $E$. For simplicity we consider piecewise constant profiles of SOCs, and solve the DW equation in each region of constant couplings. Then we introduce the $x$-dependent $4\times 4$ matrix $\Omega(x)$, whose columns are given by the components of the independent, normalized eigenspinor of the 1D DW Hamiltonian at fixed energy.\cite{footnote1} Due to the continuity of the wave function at each interface between regions of different SOC, the wave function on the left of the interface can be expressed in terms of the wave function on the right via the transfer matrix \begin{equation}\label{eq:tran:mat} \mathcal{M}= \left[\Omega(x_0^-)\right]^{-1}\Omega(x_0^+), \end{equation} where $x_0$ is the position of the interface and $x_0^\pm=x_0\pm \delta$ with infinitesimal positive $\delta$. The condition $\det \mathcal{M}=1$ guarantees conservation of the probability current across the interface. The generalization to the case of a sequence of $N$ interfaces at positions $x_i$, $i=1,\dots,N$, is straightforward since the transfer matrices relative to individual interfaces combine via matrix multiplication: \begin{equation}\label{eq:tran:mat:2} \mathcal{M} = \prod_{i=1}^N \left[\Omega(x_i^-)\right]^{-1}\Omega(x_i^+). \end{equation} From the transfer matrix it is straightforward to determine transmission and reflection matrices, which encode all the relevant information on the scattering properties. \section{The N-SO interface} \label{sec:3} \begin{figure}[!tb] \begin{center} \includegraphics[width=\columnwidth]{figure2} \caption{\label{fig:angles} (Color online) Refraction angles as function of the incidence angle for fixed energy and fixed SOCs. Panel (a): $E=5$, $\lambda=0.5$, $\Delta=2$; panel (b): $E=5$, $\lambda=2$, $\Delta=0.5$.} \end{center} \end{figure} First we concentrate on the elastic scattering problem at the interface separating a normal region N ($x<0$), where SOCs vanish, and a SO region ($x>0$), where SOCs are finite and uniform. We consider a quasiparticle of energy $E$, with $E$ assumed positive for definiteness and outside the gap possibly opened by SOCs. This quasiparticle incident from the normal region is characterized by the $y$-component of the momentum, or equivalently, the incidence angle $\phi$ measured with respect to the normal at the interface, see Fig.~\ref{fig:kinetics}. Conservation of $k_y$ implies that \begin{subequations} \begin{align} k_y^\text{N} &= E \sin \phi = E_{\alpha}\sin \xi_\alpha = k_y^\text{SO} \label{eq:mc} \\ k_x^\text{N} & = E \cos \phi \label{eq:mc:a}\\ k_{x\alpha}^\text{SO} & = E_{\alpha} \cos\xi_{\alpha}\label{eq:mc:b} \end{align} \end{subequations} where $\alpha=\pm 1$ and $E_\alpha=\sqrt{(E-\Delta)(E+\Delta-\alpha\lambda)}$. The refraction angles $\xi_\alpha$ are fixed by momentum conservation along the interface (\ref{eq:mc}) and read \begin{equation}\label{eq:ref:angles} \xi_\alpha = \arcsin\left( \frac{E}{E_\alpha}\sin\phi \right)\,. \end{equation} Figure~\ref{fig:kinetics} illustrates the refraction process at the N-SO interface. The incident wave function, assumed to have fixed spin projection in the $z$-direction, in the SO region splits in a superposition of eigenstates of the SOCs Hamiltonian corresponding to states in the different branches of the spectrum. These eigenstates propagate along two distinct directions characterized by the angles $\xi_\alpha$, whose difference depends on SOC and is an increasing function of the incidence angle, see Fig.~\ref{fig:angles}. The angles $\xi_\alpha$ coincide only for normal incidence or for $\lambda=0$. Equation (\ref{eq:ref:angles}) implies that there exists a critical angle for each of the two modes given by \begin{equation}\label{eq:crit:angles} \tilde\phi_\alpha = \arcsin\left( \frac{E_\alpha}{E} \right). \end{equation} For $\phi$ larger than both critical angles $\tilde \phi_\alpha$, the quasiparticle is fully reflected, since there are no available transmission channels in the SO region. For $\phi$ in between the two critical angles the quasiparticle transmits only in one channel.\cite{footnote2} After this qualitative discussion of the kinematics, we now present the exact solution of the scattering problem. In the N region $x<0$ a normalized scattering state of energy $E>0$, incident from the left on the interface with incidence angle $\phi$ and spin projection $s$ is given by \begin{align}\label{eq:wf:norm} \chi_\text{N}(x) = & \left[ \delta_{\uparrow,s} \kets{\uparrow} + \delta_{\downarrow,s} \kets{\downarrow}\right] \begin{pmatrix} 1\\ \text{e}^{\text{i} \phi} \end{pmatrix} \frac{\text{e}^{\text{i} k_x x}}{\sqrt{ 2v_\text{F}^x}} \nonumber\\ & + \left[ r_{\uparrow s} \kets{\uparrow}+ r_{\downarrow s} \kets{\downarrow} \right] \begin{pmatrix} 1\\ -\text{e}^{-\text{i}\phi} \end{pmatrix} \frac{\text{e}^{-\text{i} k_x x}}{\sqrt{ 2v_\text{F}^x}}, \end{align} where $k_x\equiv k_x^\text{N}$ (cf. Eq.~\ref{eq:mc:a}). Here the index $s=\uparrow,\downarrow$ specifies the spin projection of the incoming quasiparticle with $\kets{\!\uparrow}$ and $\kets{\!\downarrow}$ eigenstates of $s_z$ and $\delta_{i,j}$ is the Kronecker delta. The velocity $v_\text{F}^x=\cos\phi$ is included to ensure proper normalization of the scattering state. The complex coefficients $r_{s' s}$ are reflection probability amplitudes for a quasiparticle with spin $s$ to be reflected with spin $s'$. The associated matrix $\Omega_\text{N}(x)$ reads \begin{align} \Omega_\text{N}(x) & = \frac{1}{\sqrt{ 2v_\text{F}^x}} \nonumber \\ & \begin{pmatrix} \text{e}^{\text{i} k_x x} & \text{e}^{-\text{i} k_x x} & 0 & 0\\ \text{e}^{\text{i} (k_x x+\phi)} & -\text{e}^{-\text{i} (k_x x+\phi)} & 0 & 0\\ 0 & 0 &\text{e}^{\text{i} k_x x} & \text{e}^{-\text{i} k_x x} \\ 0 & 0 &\text{e}^{\text{i} (k_x x+\phi)} & -\text{e}^{-\text{i} (k_x x+\phi)} \nonumber \end{pmatrix}. \end{align} Similarly the wave function in the SO region ($x>0$) can be expressed in general form as \begin{align}\label{eq:wf:so} \chi_\text{SO}(x) & = \frac{1}{\sqrt{v_{++}^x}}\left[t_+\psi_{++}(x) + r_+\bar\psi_{++}(x)\right] \nonumber \\ & + \frac{1}{\sqrt{v_{-+}^x}}\left[t_-\psi_{-+}(x) + r_-\bar\psi_{-+}(x)\right] \end{align} where $t_\pm$ (resp. $r_\pm$) are complex amplitudes for right-moving (resp. left-moving) states. The coefficient $t_\alpha$ represents the transmission amplitude into mode $\alpha$. The wave functions $\psi_{\alpha+}$ and the Fermi velocities $v_{\alpha+}^x$ in the SO region are obtained from the expressions given in App.~\ref{app:so} with the replacement $k_x\to k_{x\alpha}^\text{SO}$, where for notational simplicity the label $\text{SO}$ will be understood. The wave functions $\bar\psi_{\alpha+}$ are in turn obtained from $\psi_{\alpha+}$ by replacing $k_{x\alpha} \to -k_{x\alpha}$. The matrix $\Omega_\text{SO}(x)$ then reads \begin{figure} \includegraphics[width=\columnwidth]{figure3} \caption{\label{fig:sb:tra} (Color online) Angular dependence of the transmission probabilities $T_{+\uparrow}$ (blue dashed line) and $T_{-\uparrow}$ (red solid line) at energy $E=2.5$. The SOC are fixed as follows: (a) $\lambda=0.1$ and $\Delta=0$, (b) $\lambda=0$ and $\Delta=0.1$, and (c) $\lambda=0.5$ and $\Delta=0.1$. } \end{figure} \begin{align} \Omega_\text{SO}(x) & = \\ & \hspace{-1.5cm}\begin{pmatrix} \text{e}^{-\text{i} \xi_+ -\frac{\theta_+}{2}} & - \text{e}^{\text{i} \xi_+ -\frac{\theta_+}{2}} & \text{e}^{-\text{i} \xi_- -\frac{\theta_-}{2}}& - \text{e}^{\text{i} \xi_- -\frac{\theta_-}{2}} \\ \text{e}^{\frac{\theta_+}{2}} & \text{e}^{\frac{\theta_+}{2}} & \text{e}^{\frac{\theta_-}{2}}& \text{e}^{\frac{\theta_-}{2}} \\ \text{i} \text{e}^{\frac{\theta_+}{2}} & \text{i} \text{e}^{\frac{\theta_+}{2}}& -\text{i} \text{e}^{\frac{\theta_-}{2}}& - \text{i} \text{e}^{\frac{\theta_-}{2}} \\ \text{i} \text{e}^{\text{i} \xi_+ -\frac{\theta_+}{2}} & - \text{i} \text{e}^{-\text{i} \xi_+ -\frac{\theta_+}{2}} & - \text{i} \text{e}^{\text{i} \xi_- -\frac{\theta_-}{2}} & \text{i} \text{e}^{-\text{i} \xi_- -\frac{\theta_-}{2}} \end{pmatrix} \nonumber \\ & \begin{pmatrix} \mathcal{N}_+ \text{e}^{\text{i} k_{x+}+ x} & 0 & 0 & 0 \\ 0 & \mathcal{N}_+ \text{e}^{-\text{i} k_{x+} x} & 0 & 0 \\ 0 & 0 & \mathcal{N}_- \text{e}^{\text{i} k_{x-} x} & 0 \\ 0 & 0 & 0 & \mathcal{N}_- \text{e}^{-\text{i} k_{x-} x} \end{pmatrix}\nonumber \end{align} where in the second matrix the normalization factors are defined as $\mathcal{N}_\alpha= 1/(2\sqrt{v_{\alpha+}\sinh\theta_\alpha})$. According to Eq.~(\ref{eq:tran:mat}) the transfer matrix for the single interface problem is given by the matrix product $\mathcal{M}=[\Omega_\text{N}(0^-)]^{-1}\Omega_\text{SO}(0^+)$. From $\mathcal{M}$ we obtain the transmission and the reflection probabilities for a spin-up or spin-down incident quasiparticle: \begin{align} T_{+ s} & = \displaystyle\left|\frac{\mathcal{M}_{33}\delta_{s,\uparrow} + \mathcal{M}_{13}\delta_{s,\downarrow}}{\mathcal{M}_{13}\mathcal{M}_{31}- \mathcal{M}_{11}\mathcal{M}_{33}}\right|^2 \Upsilon_+(\phi), \\ T_{- s} & = \displaystyle\left|\frac{\mathcal{M}_{31}\delta_{s,\uparrow}+ \mathcal{M}_{11}\delta_{s,\downarrow}}{\mathcal{M}_{13}\mathcal{M}_{31}- \mathcal{M}_{11}\mathcal{M}_{33}}\right|^2 \Upsilon_-(\phi),\\ R_{\uparrow s} & = \displaystyle\left|\frac{\mathcal{M}_{31}\mathcal{M}_{23}- \mathcal{M}_{33}\mathcal{M}_{21}}{\mathcal{M}_{13}\mathcal{M}_{31}- \mathcal{M}_{11}\mathcal{M}_{33}}\right|^2 \delta_{s,\uparrow} \nonumber \\ & \hspace{1.cm} + \left|\frac{\mathcal{M}_{13}\mathcal{M}_{21}-\mathcal{M}_{11}\mathcal{M}_{23}}{\mathcal{M}_{13}\mathcal{M}_{31}- \mathcal{M}_{11}\mathcal{M}_{33}}\right|^2 \delta_{s,\downarrow}, \\ R_{\downarrow s} & = \displaystyle \left|\frac{\mathcal{M}_{31}\mathcal{M}_{43}- \mathcal{M}_{33}\mathcal{M}_{41}}{\mathcal{M}_{13}\mathcal{M}_{31}-\mathcal{M}_{11} \mathcal{M}_{33}} \right|^2 \delta_{s,\uparrow} \nonumber \\ & \hspace{1cm} +\left|\frac{\mathcal{M}_{13}\mathcal{M}_{41}-\mathcal{M}_{11} \mathcal{M}_{43}}{\mathcal{M}_{13}\mathcal{M}_{31}-\mathcal{M}_{11} \mathcal{M}_{33}}\right|^2 \delta_{s,\downarrow}, \end{align} where $\Upsilon_\alpha(\phi)=\theta(\tilde\phi_\alpha-\phi)\theta(\tilde\phi_\alpha+\phi)$ with $\theta(x)$ the Heaviside step function. Here, $T_{\alpha s}$ is the probability for an incident quasiparticle with spin projection $s$ to be transmitted in mode $\alpha$ in the SO region. Of course, probability current conservation enforces $T_{+s}+T_{-s}+R_{\uparrow s}+R_{\downarrow s}=1$. Figures~\ref{fig:sb:tra} (a)--(c) show the angular dependence of the transmission probabilities for an incident spin-up quasiparticle into the $(+)$ and $(-)$ modes of the SO region for different values of the SOCs. Panel (a) refers to the case of vanishing intrinsic SOC ($\Delta=0$). Here the $(+)$ and the $(-)$ energy bands are separated by a SOC-induced splitting $\Delta E_\text{ext}=\lambda$. Therefore at fixed energy the two propagating modes in the SO region have two different momenta, which gives rise to the two different critical angles (cf. Eq.~(\ref{eq:crit:angles}) with $\Delta=0$). Panel (b) refers to the case $\lambda=0$, where the SOC opens a gap $\Delta E_\text{int}=2\Delta$ between the particle- and the hole-branches, however the $(+)/(-)$-modes remain degenerate. Therefore at fixed energy these modes have the same momentum and, as a consequence, the same critical angles (cf. Eq.~(\ref{eq:crit:angles}) for $\lambda=0$ and $\Delta\neq0$). When both SOCs are finite | the situation illustrated in panel (c) | the transmission probabilities exhibit more structure. For incidence angles smaller than $\tilde \phi_+$ no particular differences with the cases of panels (a) and (b) are visible. When the $(+)$ mode is closed, an increase (resp. decrease) of the $(-)$ mode transmission is observed for positive (resp. negative) angles, before the transmission drops to zero for incidence angles approaching $\tilde \phi_-$. The asymmetry between positive and negative angles is reversed if the spin state of the incident quasiparticle is reversed. These symmetry properties can be rationalized by considering the operator of mirror symmetry through the $x$-axes.\cite{zhai:2005} This consists of the transformation $y\to-y$ and at the same time the inversion of the spin and the pseudo-spin states. It reads \begin{equation}\label{mirror:y} \mathcal{S}_y = ( \sigma_x \otimes s_y ) R_y , \end{equation} where $R_y$ transforms $y\to-y$. The operator $\mathcal{S}_y$ commutes with the total Hamiltonian of the system $[\mathcal{S}_y, \mathcal{H}_0+\mathcal{H}_\text{SO}] =0$, therefore allows for a common basis of eigenstates. For the scattering states in the SO region (\ref{eq:wf:so}) we have $\mathcal{S}_y \chi_+(\xi_+)= \chi_+(\xi_+)$ and $\mathcal{S}_y \chi_-(\xi_-)= -\chi_-(\xi_-)$. Instead, it induces the following transformation on the scattering states (\ref{eq:wf:norm}) in the normal region: $\mathcal{S}_y \chi_s(\phi) = \text{i} \chi_{-s}(-\phi)$. By comparing the original scattering matrix with the $\mathcal{S}_y$-transformed one we find that \begin{equation}\label{eq:sing:sym} T_{\alpha,s}(\phi)=T_{\alpha,-s}(-\phi) \end{equation} with $\alpha=\pm$ and $s=\uparrow,\downarrow$, which is indeed the symmetry observed in the plots. The asymmetry of the transmission coefficients occurs only when both SOCs are finite. \subsection{Edge states at the interface} \begin{figure}[b] \centering \includegraphics[width=0.7\columnwidth]{figure4} \caption{\label{fig:bs} (Color online). Energy dispersion of the edge state at the N-SO interface as a function of the momentum along the interface $k_y$ for different values of SOCs. Solution of the transcendental equation is allowed only for $|k_y|>|E|$ (white area). In all three cases shown $\eta>1/2$: $\Delta=1$ and $\lambda=0.4$ (lower-red line), $\Delta=1.5$ and $\lambda=0.7$ (middle-blue line), and $\Delta=2$ and $\lambda=0.9$ (upper-green line).} \end{figure} In addition to scattering solutions of the DW equation, it is interesting to study the possibility that edge states exist at the N-SO interface, which propagate {\em along} the interface but decay exponentially away from it. The interest in these types of solutions is connected to the study of topological insulators. It has been shown | first by Kane and Mele\cite{kane:2005} | that a zig-zag graphene nanoribbon with intrinsic SOC supports dissipationless edge states within the SOC gap. In fact, similar states are always expected to exist at the interface between a topologically trivial and a topologically non-trivial insulator. In our case, the latter is represented by graphene with intrinsic SOC. Of course SOC-free graphene is not an insulator, however it is topologically trivial and edge state solutions do arise for $|k_y|>|E|$. When $E$ is within the gap in the SO region the corresponding mode is evanescent along the $x$ direction on both sides of the interface. Note that the edge state we find is different from the one discussed in Refs.~\onlinecite{kane:2005}, \onlinecite{stauber:2009} where zig-zag or hard-wall boundary conditions at the edge of the SOC region were imposed. The wave function on the N side then reads \begin{align} \chi_\text{N}(x) = \begin{pmatrix} 1 \\ \frac{-\text{i} q+\text{i} k_y}{E} \end{pmatrix} \left( A \kets{\downarrow} + B\kets{\uparrow} \right) \text{e}^{qx} \end{align} with $q=\sqrt{|k_y|^2-E^2}$. In the SO region the wave function can be written as \begin{align} \chi_\text{SO}(x) & = C \begin{pmatrix} (- q_+ + k_y)\\ \text{i}(E-\Delta)\\ E-\Delta\\ \text{i}(q_+ + k_y) \end{pmatrix} \text{e}^{-q_+x} + D \begin{pmatrix} (q_- - k_y)\\ -\text{i}(E-\Delta)\\ E-\Delta\\ (q_- + \text{i} k_y) \end{pmatrix} \text{e}^{-q_-x}\nonumber \end{align} with $q_\alpha=\sqrt{k_y^2-(E-\Delta)(E+\Delta-\alpha \lambda)}$. The continuity of the wave function at the N-SO interface leads to a linear system of equations for the amplitudes $A$ to $D$. The matrix of coefficients must have a vanishing determinant for a non-trivial solution to exist. This condition provides a transcendental equation for the energy of possible edge states, whose solutions are illustrated in Fig.~\ref{fig:bs}\, for different values of the intrinsic and extrinsic SOCs. The condition $|k_y|>|E|$ implies that solutions only exist outside the shadowed area. In addition, they are allowed only in the case SOCs open a gap in the energy spectrum, which occurs when $\eta>1/2$ (see App.~\ref{app:so} and Eq.~(\ref{eq:eta})). As can be seen in Fig.~\ref{fig:bs} the result is quite insensitive to the precise value of the extrinsic SOC. Edge states exist only for values of the momentum along the interface larger than the intrinsic SOC, \emph{i.e.}, $k_y>k_y^\text{min}=\Delta$. The apparent breaking of time-reversal invariance (the dispersion is not even in $k_y$) is due to the fact that we are considering a single-valley theory. The full two-valley SOC Hamiltonian is invariant under time-reversal symmetry, that interchanges the valley quantum number. This invariance implies that solutions for negative values of $k_y$ can be obtained by considering the Dirac-Weyl Hamiltonian relative to the other valley. The two counter-propagating edge states live then at opposite valleys and have opposite spin state and realize a peculiar 1D electronic system. \begin{figure} \centering \includegraphics[width=\columnwidth]{figure5} \caption{\label{figure:d}(Color online). Panel (a): Angular plots for $T_{\uparrow\uparrow}$ as a function of the injection angle for $E=2$, $\Delta=1$ and $\lambda=0$. The three lines correspond to different distance between the interfaces: $d=\pi/2$ (dashed black), $d=\pi$ (dotted red), and $d=2\pi$ (solid blue). The spin-precession length is $\ell_\text{SO}=\pi$. When $\lambda=0$ the transmission probability in the spin state opposed to the injected spin is always zero. Panel (b) and (c): angular plots of $T_{\uparrow\uparrow}$ (solid-blue) and $T_{\downarrow\uparrow}$ (dashed red) as a function of the injection angle for $E=2$, $\lambda=1$ and $\Delta=0$. The distance between the two interfaces is $d=\pi$ in panel (a) and $d=2\pi$ in panel (b). The spin-precession length is $\ell_\text{SO}=2\pi$.} \end{figure} As mentioned in the Introduction, the intrinsic SOC is estimated to be much smaller than the extrinsic one, therefore in a realistic situation one would not expect the opening of a significant energy gap and the presence of edge states. It would be interesting to explore the possibility to artificially enhance the intrinsic SOC, thereby realizing the condition for the occurrence of edge states. \section{The N-SO-N interface} \label{NSON} The analysis of the scattering problem on a N-SO interface of the previous section can be straightforwardly generalized to the case of a double N-SO-N interface (SO barrier). Here the transmission matrix $\mathcal{D}$ is given by Eq.~(\ref{eq:tran:mat:2}) with $N=2$. The transmission and the reflection probabilities in the case of a spin-up or -down incident quasiparticle read \begin{align}\label{eq:tf:di} T_{\uparrow s} & = \displaystyle\left|\frac{\mathcal{D}_{33}\delta_{s,\uparrow} +\mathcal{D}_{13}\delta_{s,\downarrow}}{\mathcal{D}_{13}\mathcal{D}_{31}-\mathcal{D}_{11}\mathcal{D}_{33}}\right|^2, \\ T_{\downarrow s} & = \displaystyle\left|\frac{\mathcal{D}_{31}\delta_{s,\uparrow}+\mathcal{D}_{11}\delta_{s,\downarrow}}{\mathcal{D}_{13}\mathcal{D}_{31}-\mathcal{D}_{11}\mathcal{D}_{33}}\right|^2, \\ R_{\uparrow s} & = \displaystyle\left|\frac{\mathcal{D}_{31}\mathcal{D}_{23}-\mathcal{D}_{33}\mathcal{D}_{21}}{\mathcal{D}_{13}\mathcal{D}_{31}-\mathcal{D}_{11}\mathcal{D}_{33}}\right|^2 \delta_{s,\uparrow} \nonumber \\ & \hspace{1.cm} + \left|\frac{\mathcal{D}_{13}\mathcal{D}_{21}-\mathcal{D}_{11}\mathcal{D}_{23}}{\mathcal{D}_{13}\mathcal{D}_{31}-\mathcal{D}_{11}\mathcal{D}_{33}}\right|^2 \delta_{s,\downarrow}, \\ R_{\downarrow s} & = \displaystyle \left|\frac{\mathcal{D}_{31}\mathcal{D}_{43}-\mathcal{D}_{33}\mathcal{D}_{41}}{\mathcal{D}_{13}\mathcal{D}_{31}-\mathcal{D}_{11}\mathcal{D}_{33}} \right|^2 \delta_{s,\uparrow} \nonumber \\ & \hspace{1cm} + \left|\frac{\mathcal{D}_{13}\mathcal{D}_{41}-\mathcal{D}_{11}\mathcal{D}_{43}}{\mathcal{D}_{13}\mathcal{D}_{31}-\mathcal{D}_{11}\mathcal{D}_{33}}\right|^2 \delta_{s,\downarrow}\, . \end{align} In this case there is an additional parameter which controls the scattering properties of the structure, namely the width $d$ of the SO region. In order to compare this length to a characteristic length scale of the system, we introduce the spin-precession length defined as \begin{equation}\label{eq:spin:pre:len} \ell_\text{SO} = 2\pi \frac{\hbar v_\text{F}}{\lambda+2\Delta}\,. \end{equation} The intrinsic SOC alone cannot induce a spin precession on the carriers injected into the SO barrier | an injected spin state, say up, is obviously never converted into a spin-down state. Figu\-re~\ref{figure:d}(a) shows the angular dependence of the transmission in the case of injection of spin-up. The behavior of the transmission as a function of the injection angle depends sensitively on the width $d$ compared to the spin-precession length. For small width $d<\ell_\text{SO}$ (dashed line) the transmission is a smooth decreasing function of $\phi$ and stays finite also for $\phi$ larger than the critical angle. In the case $d\ge\ell_\text{SO}$ (dotted- and solid-lines) instead the transmission probability exhibits a resonant behavior and drops to zero as soon as the injection angle equals the critical angle. \begin{figure} \centering \includegraphics[width=\columnwidth]{figure6} \caption{\label{figure:f}(Color online). Angular plot of $T_{\uparrow\uparrow}$ (solid-blue) and $T_{\downarrow\uparrow}$ (dashed-red) as a function of the injection angle for $E=2$, $\lambda=1$ and (a) $\Delta=\lambda/4$, (b) $\Delta=\lambda/2$, and $\Delta=\lambda$. The distance between the two interfaces is kept fixed to $d=\ell_\text{SO}$.} \end{figure} When only the extrinsic SOC is finite, the transmission behavior changes drastically. Two different critical angles appear | the biggest coincides usually with $\pi/2$. The extrinsic SOC induces spin precession because of the coupling between the pseudo- and the real-spin. This is illustrated in Fig.~\ref{figure:d}(b)-(c). In Panel (b) we consider the case of spin-up injection with $d=\ell_\text{SO}/2$. At normal incidence the transmission is entirely in the spin-down channel (dashed line). Moving away from normal incidence, the transmission in the spin-up channel (solid line) increases from zero and, after the first critical angle, the transmissions in spin-up and spin-down channels tend to coincide. In panel (c) the width of the barrier is set to $d=\ell_\text{SO}$. Here, the width of the SO region permits to an injected carrier at normal incidence to perform a complete precession of its spin state | the transmission is in the spin-up channel. For finite injection angles the spin-down transmission (dashed line) also becomes finite. For $\phi\lesssim\tilde\phi_+$ the transmission in the spin-up channel is almost fully suppressed while that in the spin-down channel is large. Finally, for $\phi>\tilde\phi_+$ the two transmission coefficients do not show appreciable difference. \begin{figure} \centering \includegraphics[width=0.7\columnwidth]{figure7a} \includegraphics[width=0.6\columnwidth]{figure7b} \caption{\label{figure:g} (Color online). Panel (a): total transmission $T$ as a function of the injection angle for $E=2$, $d=2\pi$ and several values of SOCs: $\lambda=1$ and $\Delta=0$ (blue-solid line), $\lambda=0$ and $\Delta=0.5$ (red-dotted line), $\lambda=1$ and $\Delta=\lambda/4$ (yellow-dashed line), $\Delta=\lambda/2$ (orange-dashed-dotted line), and $\lambda=\Delta$ (black-dotted-dotted-dashed line). Panel (b): $z$-component of the spin polarization ${\mathcal P}_z$ as a function on the injection angle for $E=2$ and $d=2\pi$ and the following values of the SOCs: $\lambda=1$, $\Delta=0$ and $\lambda=0$, $\Delta=1$ (same black-dashed line), $\lambda=1$ and $\Delta=\lambda/4$ (red-dotted), $\Delta=\lambda/2$ (blue-dotted-dashed line), and $\Delta=\lambda$ (green-solid line).} \end{figure} In the case where both extrinsic and intrinsic SOC are finite, the transmission probability exhibits a richer structure. We focus again on the case of injection of spin-up quasiparticles. Moreover we fix the width of the SO region so that it is always equal to the spin-precession length $d=\ell_\text{SO}$. Fig.~\ref{figure:f} illustrates the transmission probabilities $T_{s\uparrow}$ for three values of the ratio $\Delta/\lambda=1/4,1/2,1$. Notice that from panel (a) to (c) the width of SO region decreases. The symmetry properties of the transmission function can be rationalized by using the symmetry operation~(\ref{mirror:y}). Proceeding in a similar manner as in the case of the single interface, for the SO barrier we find the following symmetry relations \begin{subequations}\label{eq:sym:db} \begin{align} T_{s,s}(\phi) & = T_{s,s}(-\phi)\,, \\ T_{s,-s}(\phi) & = T_{-s,s}(-\phi)\, , \end{align} \end{subequations} which are confirmed by the explicit calculations. So far we have considered the injection of a pure spin state | the injected carrier was either in a spin-up state or a spin-down state. Following Ref. \onlinecite{marigliano2} we now address the transmission of an unpolarized statistical mixture of spin-up and spin-down carriers. This will characterize the spin-filtering properties of the SO barrier. In the injection N region, an unpolarized statistical mixture of spins is defined by the density matrix \begin{equation}\label{eq:rho:in} \rho_\text{in} = \frac{1}{2} \ket{\chi_\uparrow}\bra{\chi_\uparrow} + \frac{1}{2} \ket{\chi_\downarrow}\bra{\chi_\downarrow} , \end{equation} where $\ket{\chi_s}\equiv \ket{s}\otimes \ket{\sigma}$ with $\ket{\sigma}= (1/\sqrt{2})(1,\text{e}^{\text{i} \phi})$ corresponds to a pure spin state. When traveling through the SO region, the injected spin-unpolarized state is subjected to spin-precession. The density matrix in the output N region can be expressed in terms of the transmission functions~(\ref{eq:tf:di}) as \begin{equation}\label{eq:rho:out} \rho_\text{out} = \frac{1}{2} T_\uparrow \ket{\zeta_\uparrow}\bra{\zeta_\uparrow} + \frac{1}{2} T_\downarrow \ket{\zeta_\downarrow}\bra{\zeta_\downarrow}, \end{equation} where the coefficients $T_s=T_{\uparrow s}+T_{\downarrow s}$ are the total transmissions for fixed injection state. The spinor part is defined as \begin{equation}\label{eq:spinor:sm} \ket{\zeta_s}=\frac{1}{\sqrt{T_s}} \begin{pmatrix} t_{\uparrow s} \\ t_{\downarrow s} \end{pmatrix} \otimes \ket{\sigma}, \end{equation} where $t_{s',s}$ are the transmission amplitudes for incoming (resp. outgoing) spin $s$ (resp. $s'$). The output density matrix is used to define the total transmission \begin{equation}\label{eq:tran:unp} T=\frac{T_\uparrow +T_\downarrow}{2} \end{equation} and the expectation value of the $z$ component of the spin-polarization \begin{equation}\label{eq:pol:z} \mathcal{P}_z = \frac{1}{2} \left( T_{\uparrow\uparrow} + T_{\uparrow\downarrow} - T_{\downarrow\uparrow} - T_{\downarrow\downarrow}\right)\,. \end{equation} In Fig.~\ref{figure:g} we report the total transmission (panel~(a)) and the $z$-component of the spin-polarization (panel~(b)) as a function of the injection angle for fixed energy and width of the SO region. We observe that for an unpolarized injected state the transmission probability is an even function of the injection angle $T(\phi)=T(-\phi)$. Moreover, for injection angles larger than the first critical angle $\phi > \tilde\phi_+$, the transmission has an upper bound at $T = 1/2$. On the contrary $\mathcal{P}_z$ is an odd function of the injection angle $\mathcal{P}_z(\phi)=-\mathcal{P}_z(-\phi)$. It is zero when at least one SOC is zero. When both SOC parameters are finite $\mathcal{P}_z$ is finite and reaches the largest values for $\phi > \tilde\phi_+$. The maxima in this case increase as a function of the intrinsic SOC. To experimentally observe this polarization effect the measurement should not involve an average over the angle $\phi$, which, otherwise | due to the antisymmetry of $\mathcal{P}_z$ | would wash out the effect. To achieve this, one could use, {\it e.g.}, magnetic barriers,\cite{ale,luca1} which are known to act as wave vector filters. \section{Conclusions} \label{sec:5} In this paper we have studied the spin-resolved transmission through SO nanostructures in graphene, {\em i.e.}, systems where the strength of SOCs | both intrinsic and extrinsic | is spatially modulated. We have considered the case of an interface separating a normal region from a SO region, and a barrier geometry with a region of finite SOC sandwiched between two normal regions. We have shown that | because of the lift of spin degeneracy due to the SOCs | the scattering at the single interface gives rise to spin-double refraction: a carrier injected from the normal region propagates into the SO region along two different directions as a superposition of the two available channels. The transmission into each of the two channels depends sensitively on the injection angle and on the values of SOC parameters. In the case of a SO barrier, this result can be used to select preferential directions along which the spin polarization of an initially unpolarized carrier is strongly enhanced. We have also analyzed the edge states occurring in the single interface problem in an appropriate range of parameters. These states exist when the SOCs open a gap in the energy spectrum and correspond to the gapless edge states supported by the boundary of topological insulators. A natural follow-up to this work would be the detailed analysis of transport properties of such SO nanostructures. From our results for the transmission probabilities, spin-resolved conductance and noise could easily be calculated by means of the Landauer-B\"uttiker formalism. Moreover we plan to study other geometries, as, \emph{e.g.}, nanostructures with a periodic modulation of SOCs. The effects of various types of impurities on the properties discussed here is yet another interesting issue to address. We hope that our work will stimulate further theoretical and experimental investigations on spin transport properties in graphene nanostructures. \acknowledgments We gratefully acknowledge helpful discussions with L.~Dell'Anna, R.~Egger, H.~Grabert, M.~Grifoni, W. H\"ausler, V.~M.~Ramaglia, P.~Recher and D.~F.~Urban. The work of DB is supported by the Excellence Initiative of the German Federal and State Governments. The work of ADM is supported by the SFB/TR 12 of the DFG.
\section{Introduction } In this article we consider the dynamics of a collection of \textit{hard Brownian spheres} with drifts or boundary conditions that includes instantaneous reflections upon collisions. The models are similar to existing ones in the literature that consider point masses instead of spheres of a positive radius. We will show that the (empirical) distribution of a family of Brownian spheres behaves differently from the (empirical) distribution of the point Brownian particles in some natural models. In particular, the distribution of Brownian spheres fails to satisfy the usual heat equation under circumstances that lead to the heat equation for the infinitely many infinitesimally small Brownian particles. Various models of colliding Brownian particles have been considered in the statistical physics literature. One stream, pioneered by \cite{H}, considers a countable collection of Brownian point masses on the line that collide and reflect instantaneously. Also see the follow-up work on tagged particle in the Harris model by D\"urr, Goldstein, \& Lebowitz, \cite{DGL}. A variation on the theme has been to replace the instantaneous reflection by a potential, and goes by the name of \textit{gradient systems}. In these gradient systems, one studies the behavior of countably many particles under a repelling potential. Usually the potential is modeled as smooth with a singularity at zero; see the article by \cite{CL}. A particular example of this class includes the famous Dyson Brownian motion from Random Matrix theory; see \cite{D62}, and \cite{CL2}. The other class of models, closer to our article, goes by the name of \textit{hard-core} interactions, in which the Brownian particles are assumed to be hard balls of small radius, and consequently, there is instantaneous reflection whenever two such balls collide (plus possible additional interactions). This is the spirit taken in the articles by Dobrushin \& Fritz in dimension one, \cite{DF}, and Fritz \& Dobrushin in dimension two, \cite{FD}, \cite{L77,L78} (with a correction by \citealp{S79}). The main focus of these authors is the non-equilibrium dynamics of the gradient systems. Also see the articles by \cite{O96,O98}, and \cite{T96} all of which consider properties of a tagged particle in the infinite system. In the discrete case, the various models of symmetric and asymmetric exclusion processes have been considered. Closest in spirit to the models discussed here is the totally asymmetric exclusion process (TASEP) considered by Baik, Deift, \& Johansson, \cite{BDJ}, and \cite{J00} in connection with random matrices and the longest increasing subsequence problem. Specifically, if the initial configuration in TASEP is $\mathbb{Z}^-$, then the probability that a particle initially at $-m$ moves at least $n$ steps to the right in time $t$ equals the probability distribution of the largest eigenvalue in a unitary Laguerre random matrix ensemble. In recent subsequent articles \cite{TW,TW09}, Tracy \& Widom explicitly compute transition probabilities of individual particles in the asymmetric exclusion process, extending Johansson's work. In this paper, we consider only one dimensional models, so our ``spheres'' are actually intervals. The title of this paper reflects our intention to study multidimensional models in future articles. We leave more detailed discussion to Section~\ref{sec:disc}. That section also contains references to related research projects. We consider two models, which have the following common features. Informally speaking, both models consist of families of Brownian ``particles''. The $k$-th ``particle'' is represented by an interval $I^k_t = (X^k_t-\varepsilon/2, X^k_t+\varepsilon/2)$, where $X^k_t$ is a Brownian-like process. The intervals $I^k$ and $I^j$ are always disjoint, for $k\ne j$. The processes $X^k$ are driven by independent Brownian motions. When two intervals $I^k$ and $I^j$ collide, they reflect instantaneously. In the first model, the number of particles is constant and they are pushed by a barrier moving at a constant speed. In the second model, particles enter the interval $[0,1]$ at the left, they reflect at 0, and they are killed when they hit the right endpoint. The second model is our primary focus because it is related to other models considered in mathematical physics literature---see Section \ref{sec:disc}. We are grateful to Thierry Bodineau, Pablo Ferrari, Claudio Landim, Mario Primicerio and Jeremy Quastel for very helpful advice. We would like to thank the referee for the suggestions for improvement. \section{Extreme crowding} We start with an informal description of our first model, which consists of a fixed number $n$ of ``particles''. The $k$-th leftmost ``particle'' is represented by an interval $I^k_t = (X^k_t-\varepsilon/2, X^k_t+\varepsilon/2)$. The intervals $I^k$ and $I^j$ are always disjoint. The processes $X^k$ are driven by independent Brownian motions. When two intervals $I^k$ and $I^j$ collide, they reflect instantaneously. The intervals are pushed from the left by a barrier with a constant velocity, that is, the leftmost interval reflects on the line $x= ct$. Formally, we define $\{X^0,X^1,\ldots,X^n\}$ to be continuous processes such that $X^0_t=-\varepsilon/2+ct$, $X^k_t-X^{k-1}_t\ge\varepsilon$ for all $k\ge1$ and all $t\ge0$, and for $k\ge 1$, \[ dX^k_t = dB^k_t + dL^k_t - dM^k_t, \] where $\{B^1,\ldots,B^n\}$ are iid Brownian motions, and $L^k$ and $M^k$ are nondecreasing processes such that \[ \int_0^\infty 1_{\{X^k_t-X^{k-1}_t>\varepsilon\}}\,dL^k_t = 0 \qquad\text{and}\qquad \int_0^\infty 1_{\{X^{k+1}_t-X^k_t>\varepsilon\}}\,dM^k_t = 0. \] (Here, we may interpret $X^{n+1}\equiv\infty$.) The distributions of $X^k_0$ for $1\le k\le n$ will be specified later. To construct the solution to this Skorohod problem, consider first the processes $Y^k_t=X^k_t-(k-1)\varepsilon-\varepsilon/2-ct$. These processes satisfy $Y^0\equiv0$, $Y^k_t-Y^{k-1}_t\ge0$ for all $k\ge1$ and all $t\ge0$, and for $k\ge 1$, $dY^k_t = dB^k_t - c\,dt + dL^k_t - dM^k_t$, where \[ \int_0^\infty 1_{\{Y^k_t-Y^{k-1}_t>0\}}\,dL^k_t = 0 \qquad\text{and}\qquad \int_0^\infty 1_{\{Y^{k+1}_t-Y^k_t>0\}}\,dM^k_t = 0. \] We may therefore construct the processes $\{Y^1,\ldots,Y^n\}$ using order statistics.\break Namely, let $\{Z^1,\ldots,Z^n\}$ be defined by $dZ^k_t = dB^k_t - c dt$, and reflected at 0. For every fixed $t\geq 0$, we let $Y^1_t, Y^2_t, \dots, Y^n_t$ be ordered $Z^k_t$'s, that is, $\{Y^1_t, \dots, Y^n_t\} =\{Z^1_t, \dots, Z^n_t\}$ and $Y^1_t \leq Y^2_t\leq \dots \leq Y^n_t$. Finally, we let $X^k_t = Y^k_t +(k-1) \varepsilon + \varepsilon/2 + c t$ and $I^k_t = (X^k_t-\varepsilon/2, X^k_t+\varepsilon/2)$. Let $n\varepsilon = b$. We will fix $b>0$ and analyze the behavior of the system of intervals $\{I^k\}$ as $n\to\infty$. In other words, we will keep the total length of all intervals $I^k$ constant. The stationary distribution for $Z^k$ has the density $\varphi(z) = ce^{-cz}$ for $z\geq 0$, with $c= 2c_1$, because \begin{align*} \frac 1 2 \frac{d^2}{dz^2} \varphi(z) + c_1 \frac d {dz} \varphi(z) =0. \end{align*} Consider any $0\leq x_1 < x_2 < \infty$, let $\lambda$ denote the Lebesgue measure, and let \begin{align}\label{eq:m17.4} {\bf d}([x_1, x_2]) = {\bf d}_t([x_1, x_2]) = \frac { \lambda\left( [x_1 + ct, x_2 + ct] \cap \bigcup_{1\leq k \leq n} I^k_t \right)} {x_2 - x_1}. \end{align} The quantity ${\bf d}([x_1, x_2])$ represents the average density of ``particles'' $I^k$ on the interval $[x_1, x_2]$. We will say that the intervals $\{I^k\}$ have the pseudo-stationary distribution if all $Z^k_t$'s are independent and have the stationary distribution $\varphi$ for $t=0$ and, therefore, for every $t\geq 0$. \begin{theorem}\label{th:m16.1} Suppose that the intervals $\{I^k\}$ have the pseudo-stationary distribution. Fix arbitrary $p_1, d_1< 1$, $d_2>0$, and $0\leq x_1 < x_2 < b < x_3 < x_4 < \infty$. There exist $c_0, n_0 < \infty$ such that for $c\geq c_0$, $n \geq n_0$ and any $t\geq 0$, we have \begin{align}\label{eq:m17.1} P({\bf d}_t([x_1, x_2]) \geq d_1) &\geq p_1,\\ P({\bf d}_t([x_3, x_4]) \leq d_2) &\geq p_1. \label{eq:m17.2} \end{align} \end{theorem} The theorem says that the ``particles'' $I^k$ clump together and there is a sharp transition in density of ``mass'' around $x = b$. This is in contrast with infinitely small ``particles'' $Z^k$ whose empirical distribution is close to the distribution with the density $\varphi(z) = c e^{-cz}$ that displays no sharp drop-off. \begin{proof}[Proof of Theorem \ref{th:m16.1}] Without loss of generality, we let $t=0$. We define $y_k\in(0,\infty)$ in an implicit way by the following formula, for $k=1, 2,3,4$, \begin{align*} x_k = b \int_0^{y_k} \varphi(z) dz + y_k. \end{align*} Note that $y_1 < y_2$, and that for $\varepsilon>0$ sufficiently small (that is, for $n=b\varepsilon^{-1}$ sufficiently large), it is possible to choose $y_5,y_6$ such that $y_1< y_5 < y_6 < y_2$, and \begin{align}\label{eq:m16.2} \frac {b \int_{y_5}^{y_6} \varphi(z) dz- 2\varepsilon} {y_2 - y_1 + b \int_{y_1}^{y_2} \varphi(z) dz } \geq \frac {b \int_{y_1}^{y_2} \varphi(z) dz} {y_2 - y_1 + b \int_{y_1}^{y_2} \varphi(z) dz } - (1-d_1)/2. \end{align} Since $b-x_2 >0$, we can find $c$ so large that, \begin{align}\label{eq:m16.4} \frac {c (b -x_2 ) } {1 + c (b -x_2 ) } \geq 1- (1-d_1)/2. \end{align} Let $\lceil a\rceil$ denote the smallest integer greater than or equal to $a$. By the law of large numbers, if $n$ is sufficiently large, the number of $Z^k_0$'s in the interval $[0,y_1]$ is smaller than or equal to $n\int_{0}^{y_5} \varphi(z) dz$, with probability greater than $1-(1-p_1)/2$. If this event holds then there are exactly $\left\lceil n\int_{0}^{y_5} \varphi(z) dz\right\rceil$ processes $Z^k_0$ in some (random) interval $[0,y_7]$ with $y_7 \geq y_1$. This implies that there are exactly $\left\lceil n\int_{0}^{y_5} \varphi(z) dz\right\rceil$ processes $X^k_0$ in $[0, \varepsilon \left\lceil n\int_{0}^{y_5} \varphi(z) dz\right\rceil + y_7]$. Note that \begin{align*} \varepsilon \left\lceil n\int_{0}^{y_5} \varphi(z) dz\right\rceil + y_7 \geq b \int_{0}^{y_5} \varphi(z) dz + y_7 \geq b\int_{0}^{y_1} \varphi(z) dz + y_1 = x_1. \end{align*} Hence, the number of $X^k_0$'s in the interval $[0, x_1]$ is smaller than or equal to $n\int_{0}^{y_5} \varphi(z) dz+1$, with probability greater than $1-(1-p_1)/2$. A completely analogous argument shows that, if $n$ is sufficiently large, then the number of $X^k_0$'s in the interval $[x_2, \infty]$ is smaller than or equal to $n\int_{y_6}^{\infty} \varphi(z) dz+1$, with probability greater than $1-(1-p_1)/2$. Both events hold with probability greater than $1-2(1-p_1)/2 = p_1$, and then the number of $X^k_0$'s in $[x_1, x_2]$ is greater than or equal to $n\int_{y_5}^{y_6} \varphi(z) dz-2$. This and \eqref{eq:m16.2} imply that \begin{align}\label{eq:m16.3} {\bf d}([x_1,x_2]) &\geq \frac {\varepsilon n\int_{y_5}^{y_6} \varphi(z) dz-2\varepsilon} {x_2 - x_1 } = \frac {b \int_{y_5}^{y_6} \varphi(z) dz - 2\varepsilon} {b \int_0^{y_2} \varphi(z) dz + y_2 - b \int_0^{y_1} \varphi(z) dz - y_1 } \nonumber \\ &= \frac {b \int_{y_5}^{y_6} \varphi(z) dz - 2\varepsilon} {y_2 - y_1 + b \int_{y_1}^{y_2} \varphi(z) dz } \geq \frac {b \int_{y_1}^{y_2} \varphi(z) dz} {y_2 - y_1 + b \int_{y_1}^{y_2} \varphi(z) dz } - (1-d_1)/2. \end{align} We have \begin{align*} x_2 = b \int_0^{y_2} \varphi(z) dz + y_2 = b \int_0^{y_2} ce^{-cz} dz + y_2 = y_2 + b - b e^{-cy_2}, \end{align*} so $e^{-cy_2} = (y_2 -x_2 +b)/b$ and, therefore, for $z\leq y_2$, \begin{align*} \varphi(z) = ce^{-cz} \geq ce^{-c y_2} = (c/b)(y_2 -x_2 +b). \end{align*} We combine this estimate with \eqref{eq:m16.3} and \eqref{eq:m16.4} to see that, with probability greater than $p_1$, \begin{align*} {\bf d}([x_1,x_2]) &\geq \frac {b \int_{y_1}^{y_2} \varphi(z) dz} {y_2 - y_1 + b \int_{y_1}^{y_2} \varphi(z) dz } - (1-d_1)/2 \\ &\geq \frac {b \int_{y_1}^{y_2} (c/b)(y_2 -x_2 +b) dz} {y_2 - y_1 + b \int_{y_1}^{y_2} (c/b)(y_2 -x_2 +b) dz } - (1-d_1)/2\\ &= \frac {c (y_2-y_1)(y_2 -x_2 +b) } {y_2 - y_1 + c (y_2-y_1)(y_2 -x_2 +b) } - (1-d_1)/2\\ &= \frac {c (y_2 -x_2 +b) } {1 + c (y_2 -x_2 +b) } - (1-d_1)/2\\ &\geq \frac {c (b -x_2 ) } {1 + c (b -x_2 ) } - (1-d_1)/2\\ &\geq 1- (1-d_1)/2 - (1-d_1)/2 = d_1. \end{align*} This completes the proof of \eqref{eq:m17.1}. The proof of \eqref{eq:m17.2} is completely analogous. \end{proof} \section{Brownian gas under pressure} \label{sec:pressure} In this model, ``particles'' $I^k$ are confined to the interval $[0,1]$. More precisely, their centers are confined to this interval. The $k$-th leftmost ``particle'' is represented by an interval $I^k_t = (X^k_t-\varepsilon/2, X^k_t+\varepsilon/2)$. The intervals $I^k$ and $I^j$ are always disjoint. The processes $X^k$ are driven by independent Brownian motions with the diffusion coefficient $\sigma^2$. When two intervals $I^k$ and $I^j$ collide, they reflect instantaneously. The particles are added to the system at the left endpoint of $[0,1]$ at a constant rate. In other words, they are pushed in at the speed $a$, so that a new particle enters the interval every $\varepsilon/a$ units of time. As soon as $X^k$ reaches 0, it starts moving as a Brownian motion reflected at 0. The $k$-th interval is removed from the system when $X^k$ hits the right endpoint of $[0,1]$. Formally, we define $\{X^1,X^2,\ldots\}$ to be a collection of right-continuous, $[0,\infty]$-valued processes such that \begin{align} &X^k_0=-k\varepsilon+\varepsilon/2 \text{ for all $k$},\\ &\text{If $S_k=\inf\{t>0: X^k_{t-}=1-\varepsilon/2\}$},\notag\\ &\qquad\text{then $X^k_t$ is continuous on $[0,S_k)$ and $X^k_t=\infty$ for all $t>S_k$},\\ &\text{$X^k_t-X^{k+1}_t\ge\varepsilon$ for all $k\ge1$ and all $t\ge0$}, \text{ and}\\ &dX^k_t = \begin{cases} a\,dt &\text{if $t\in[0,k\varepsilon/a)$},\\ \sigma\,dB^k_t + dL^k_t - dM^k_t &\text{if $t\in[k\varepsilon/a,S_k)$}, \end{cases} \end{align} where $a$ and $\sigma$ are positive constants, $\{B^1,B^2,\ldots\}$ are iid Brownian motions, and $L^k$ and $M^k$ are nondecreasing processes such that \[ \int_{k\varepsilon/a}^{S_k} 1_{\{X^k_t-X^{k+1}_t>\varepsilon\}}\,dL^k_t = 0 \qquad\text{and}\qquad \int_{k\varepsilon/a}^{S_k} 1_{\{X^{k-1}_t-X^k_t> \varepsilon\}}\,dM^k_t = 0. \] (Here, we may interpret $X^0\equiv\infty$.) To construct the solution to this Skorohod problem, consider first the processes $Y^k_t = X^k_t + k \varepsilon - \varepsilon/2 - at$. These processes satisfy \begin{align} &Y^k_0 = 0 \text{ for all $k$},\label{sko1}\\ &\text{If $S_k=\inf\{t>0: Y^k_{t-}=1-at+(k -1)\varepsilon\}$},\notag\\ &\qquad\text{then $Y^k_t$ is continuous on $[0,S_k)$ and $Y^k_t=\infty$ for all $t>S_k$},\\ &\text{$Y^k_t-Y^{k+1}_t\ge0$ for all $k\ge1$ and all $t\ge0$}, \text{ and}\\ &dY^k_t = \begin{cases} 0 &\text{if $t\in[0,k\varepsilon/a)$},\\ \sigma\,dB^k_t - a\,dt + dL^k_t - dM^k_t &\text{if $t\in[k\varepsilon/a,S_k)$}, \label{sko4} \end{cases} \end{align} where \[ \int_{k\varepsilon/a}^{S_k} 1_{\{Y^k_t-Y^{k+1}_t>0\}}\,dL^k_t = 0 \qquad\text{and}\qquad \int_{k\varepsilon/a}^{S_k} 1_{\{Y^{k-1}_t-Y^k_t>0\}}\,dM^k_t = 0. \] Again, we shall construct the processes $\{Y^1,Y^2,\ldots\}$ using order statistics. Let $Z^k$ be a $[0,\infty)$-valued process, satisfying the SDE $dZ^k_t = \sigma dB^k_t - a dt$, and reflected at $0$. The process $Z^k_t$ is defined on the time interval $t\in[k\varepsilon/a, \infty)$, and starts at $Z^k_{k\varepsilon/a} = 0$. At any time $t \in [k\varepsilon/a, (k+1)\varepsilon/ a)$, only processes $Z^j$, $1 \leq j \leq k$, are defined. Let $\lfloor a\rfloor$ denote the greatest integer less than or equal to $a$, and \begin{align*} S_0&=0,\\ {\mathcal A}^1_t &= \{j \in \mathbb {Z} : 1\leq j \leq\lfloor ta/\varepsilon \rfloor\}, \quad t\geq 0,\\ S_1 &= \inf\{t>0: \sup_{j \in {\mathcal A}^1_t} Z^j_t \geq 1- at\},\\ {\mathcal A}^k_t &= {\mathcal A}^{k-1}_t \setminus \{m\in \mathbb {Z}: Z^m_{S_{k-1}} = 1 - at + (k-2) \varepsilon\}, \quad t\geq S_{k-1}, k\geq 2,\\ S_k &= \inf\{t>S_{k-1}: \sup_{j\in{\mathcal A}^k_t} Z^j_t \geq 1- at+(k-1)\varepsilon\}, \quad k\geq 2. \end{align*} Note that it is possible that ${\mathcal A}^k_t = \emptyset$ for some random $k$ and $t>0$. \noindent{\it Convention (C)}. For the sake of future reference, it is convenient to say that the process $Z^m$ is killed at the time $S_{k-1}$, where $\{m\} = {\mathcal A}^{k-1}_{S_{k-1}} \setminus {\mathcal A}^{k}_{S_{k-1}}$. In other words, $Z^m$ is killed when the corresponding interval $I^k$, defined below, hits the right endpoint of the interval $[0,1]$. For every $t\in[S_{k-1}, S_k)$, note that there are $\lfloor ta/\varepsilon \rfloor - (k-1)$ elements in ${\mathcal A}^k_t$. Let $Y^k_t, Y^{k+1}_t, \dots, Y^{\lfloor ta/\varepsilon \rfloor}_t$ be reverse-ordered $Z^j_t$'s, $j\in{\mathcal A}^k_t$, that is, $\{Y^k_t, \dots, Y^{\lfloor ta/\varepsilon \rfloor}_t\} =\{Z^j_t, j\in{\mathcal A}^k_t\}$ and $Y^k_t \ge Y^{k+1}_t\ge \dots \ge Y^{\lfloor ta/\varepsilon \rfloor}_t$. Let $Y^j_t=\infty$ for $j<k$ and $Y_t^j=0$ for $j>\lfloor ta/\varepsilon \rfloor$. It is elementary to check that $\{Y^1,Y^2,\ldots\}$ satisfy \eqref{sko1}-\eqref{sko4}. We may therefore define \begin{align*} X^k_t &= Y^k_t - k\varepsilon + \varepsilon/2 + at, \\ I^k_t &= (X^k_t-\varepsilon/2, X^k_t+\varepsilon/2). \end{align*} We have to modify slightly the definition \eqref{eq:m17.4} of density to match the current model. For $t\in[S_{k-1}, S_k)$, let \begin{align}\label{eq:m20.1} {\bf d}([x_1, x_2])= {\bf d}_t([x_1, x_2])=\frac { \lambda\left( [x_1, x_2] \cap \bigcup_{k \leq j \leq \lfloor ta/\varepsilon \rfloor} I^j_t \right)} {x_2 - x_1}. \end{align} \begin{theorem}\label{th:m17.5} Fix arbitrary $0<x_1<x_2<1$, $p_1<1$ and $a,\sigma,c_0>0$. There exist $t_0< \infty$ and $\varepsilon_0>0$ such that for $t\geq t_0$ and $\varepsilon \in (0,\varepsilon_0)$, \begin{align}\label{eq:m17.7} P\left(\frac {1-x_2} {1-x_2 + \sigma^2/(2a)} -c_0 \leq {\bf d}_t([x_1,x_2]) \leq \frac {1-x_1} {1-x_1 + \sigma^2/(2a)} +c_0 \right) &\geq p_1. \end{align} \end{theorem} Intuitively speaking, the theorem says that the mass density at $x\in(0,1)$ is close to $(1-x)/(1-x + \sigma^2/(2a))$, for large $t$ and small $\varepsilon$. \begin{proof}[Proof of Theorem \ref{th:m17.5}] We will use the coupling technique. Recall processes\break $Z^1, Z^2, \dots$ used in the definition of $Y^k$'s---we will use the same $Z^k$'s to construct auxiliary processes. Fix some $v_1>0$, let $\widehat S_k = \inf\{t\geq 0: Z^k_t = v_1\}$, and let $\widehat Z^k_t$ be the process $Z^k$ killed at the time $\widehat S_k$. Let $n_t$ be the number of processes $\widehat Z^k$ alive at time $t$. Let $\widehat Y^1_t, \widehat Y^2_t, \dots, \widehat Y^{n_t}_t$ be ordered $\widehat Z^j_t$'s, that is, $\{\widehat Y^1_t, \dots, \widehat Y^{n_t}_t\} = \{\widehat Z^j_t, \widehat S_j > t\}$ and $\widehat Y^1_t \leq \widehat Y^2_t\leq \dots \leq \widehat Y^{n_t}_t$. For $t\in[\widehat S_{k-1}, \widehat S_k)$ and $j=k,\dots, k + n_t-1$, we let \begin{align*} \widehat X^j_t &= \widehat Y^{n_t + k - j}_t +(n_t + k - j-1) \varepsilon + \varepsilon/2 + (t- \lfloor ta/\varepsilon \rfloor \varepsilon /a) a, \\ \widehat I^j_t &= (\widehat X^j_t-\varepsilon/2, \widehat X^j_t+\varepsilon/2). \end{align*} Every process $\widehat X^j_t$ is defined on the interval $[j\varepsilon/a,\widehat S_{j})$ and it is continuous on this interval. Although it may not be apparent from the above formulas, the processes $\widehat Y^j, \widehat X^j$ and $\widehat I^j$ are constructed from $\widehat Z^j$'s in the same way as $ Y^j, X^j$ and $ I^j$ were constructed from $ Z^j$'s. We leave the verification of this claim to the reader. We will find the Green function $G_{v_1}(v)$ of $\widehat Z^k$, i.e., the density of its occupation measure. Consider a process $V$ with values in $[-v_1, v_1]$, satisfying the SDE $d V_t = dB_t - a \operatorname{sign} ( V_t) dt$, where $B$ is Brownian motion, $ V_0 = 0$, and such that $ V$ is killed when it hits $-v_1$ or $v_1$. Note that the Green function $ G^V_{v_1}(v)$ of $ V$ is one half of $G_{v_1}(v)$ for $v>0$. The scale function $S(v)$ and the speed measure $m(v)$ for $ V$ can be calculated as follows (see \citealp{KT2}, pp. 194-195), \begin{align*} s(v) & = \exp\left(\int_0^v -(-2 a \operatorname{sign}(x)/\sigma^2) dx\right) = \exp(2 a v \operatorname{sign}(v)/\sigma^2),\\ S(v) &= \int_0^v s(x) dx = \frac{\operatorname{sign}(v) \sigma^2}{2a} \left( \exp ( 2 a v \operatorname{sign}(v)/\sigma^2) -1\right),\\ m(v) &= 1/(\sigma^2 s(v)) = (1/\sigma^2) \exp(-2 a v \operatorname{sign}(v)/\sigma^2). \end{align*} We will use formula (3.11) on page 197 of \cite{KT2}. In that formula, we take $x=0$, so $u(0) = 1/2$, by symmetry. We apply the formula to functions $g(v)$ of the form $g(v) = {\bf 1}_{[v_3, v_4]}(v)$, to conclude that for $v\in(0,v_1)$, the Green function $ G^V_{v_1}(v)$ is given by \begin{align*} G^V_{v_1}(v)& = (S(v_1) - S(v)) m(v)\\ &= \frac 1 {2a} (\exp ( 2 a v_1 /\sigma^2) - \exp ( 2 a v /\sigma^2)) \exp(-2 a v /\sigma^2) \\ &= \frac 1 {2a} (\exp ( 2 a (v_1-v) /\sigma^2) -1). \end{align*} It follows that \begin{align*} G_{v_1}(v) = 2 G^V_{v_1}(v) = \frac 1 {a} (\exp ( 2 a (v_1-v) /\sigma^2) -1). \end{align*} Define $v_0\in(0,\infty)$ by setting \begin{align*} \varphi(v) = a G_{v_0}(v) = \frac 1 {a} (\exp ( 2 a (v_0-v) /\sigma^2) -1), \end{align*} and the following condition, \begin{align}\label{eq:m21.4} 1 &= \int_0^{v_0} \varphi(v) dv + v_0 = \int _0^{v_0} (\exp ( 2 a (v_0-v) /\sigma^2) -1)dv + v_0\\ &= (\sigma^2/2a) (\exp ( 2 a v_0 /\sigma^2)-1).\nonumber \end{align} We define $y_k\in(0,\infty)$, $k=1,2$, by the following formula, \begin{align}\nonumber x_k &= \int_0^{y_k} \varphi(v) dv + y_k\\ &= \int_0^{y_k} (\exp ( 2 a (v_0-v) /\sigma^2) -1) dv + y_k \nonumber \\ & = (-(\sigma^2/2a)\exp ( 2 a (v_0-v) /\sigma^2) -v)\Big|_{v=0}^{v=y_k} + y_k \nonumber \\ &= (\sigma^2/2a)(\exp ( 2 a v_0 /\sigma^2) - \exp( 2 a (v_0-y_k) /\sigma^2) ). \label{eq:m21.2} \end{align} Choose $ y_1 < y_3 < y_4 < y_2$ and $v_1< v_0$ such that, \begin{align}\label{eq:m21.1} \frac {a\int_{y_3}^{y_4} G_{v_1}(v) dv} {y_2 - y_1 + \int_{y_1}^{y_2} \varphi(z) dz } \geq \frac { \int_{y_1}^{y_2} \varphi(z) dz} {y_2 - y_1 + \int_{y_1}^{y_2} \varphi(z) dz } - c_0. \end{align} Recall that $\lceil a\rceil$ denotes the smallest integer greater than or equal to $a$. Let $\lfloor a\rfloor$ denote the largest integer smaller than or equal to $a$. Let $c_1 = 1-p_1$ and $p_2 = 1- c_1/8$. By the law of large numbers, we can find a large $t_0$ and make $\varepsilon_0>0$ smaller, if necessary, such that if $t\geq t_0$ and $\varepsilon\in(0,\varepsilon_0)$ then with probability greater than $p_2$, the number of processes $\widehat Z^k_{t}$ in the interval $[0,y_1]$ is smaller than or equal to $(a/\varepsilon)\int_0^{y_3} G_{v_1}(v) dv$. If this event holds then there are exactly $\left\lceil (a/\varepsilon)\int_0^{y_3} G_{v_1}(v) dv\right\rceil$ processes $\widehat Z^k_0$ in some (random) interval $[0,y_5]$ with $y_5 \geq y_1$. This implies that there are exactly $\left\lceil (a/\varepsilon)\int_0^{y_3} G_{v_1}(v) dv\right\rceil$ processes $\widehat X^k_0$ in $[0, \varepsilon \left\lceil (a/\varepsilon)\int_0^{y_3} G_{v_1}(v) dv\right\rceil + y_5]$. For fixed $y_1$ and $y_3$, we make $v_1< v_0$ larger, if necessary, so that \begin{align*} \varepsilon \left\lceil (a/\varepsilon)\int_0^{y_3} G_{v_1}(v) dv \right\rceil + y_5 &\geq a \int_{0}^{y_3} G_{v_1}(v) dv + y_5 \geq a \int_{0}^{y_3} G_{v_1}(v) dv + y_1 \\ &\geq a\int_{0}^{y_1} G_{v_0}(v) dv + y_1 = \int_{0}^{y_1} \varphi(v) dv + y_1 =x_1. \end{align*} Hence, the number of $\widehat X^k_0$'s in the interval $[0, x_1]$ is smaller than or equal to $(a/\varepsilon)\int_0^{y_3} G_{v_1}(v) dv$, with probability greater than $p_2$. We can make $t_0$ larger and $\varepsilon_0>0$ smaller, if necessary, so that by the law of large numbers, if $t\geq t_0$ and $\varepsilon\in(0,\varepsilon_0)$ then with probability greater than $p_2$, the number of processes $\widehat Z^k_{t_3}$ in the interval $[0,y_2]$ is greater than or equal to $(a/\varepsilon)\int_0^{y_4} G_{v_1}(v) dv$. If this event holds then there are exactly $\left\lfloor (a/\varepsilon)\int_0^{y_4} G_{v_1}(v) dv\right\rfloor$ processes $\widehat Z^k_0$ in some (random) interval $[0,y_6]$ with $y_6 \leq y_2$. This implies that there are exactly $\left\lfloor (a/\varepsilon)\int_0^{y_4} G_{v_1}(v) dv\right\rfloor$ processes $\widehat X^k_0$ in $[0, \varepsilon \left\lceil (a/\varepsilon)\int_0^{y_4} G_{v_1}(v) dv\right\rceil + y_6]$. Note that, \begin{align*} \varepsilon \left\lfloor (a/\varepsilon)\int_0^{y_4} G_{v_1}(v) dv \right\rfloor + y_6 &\leq a \int_{0}^{y_4} G_{v_1}(v) dv + y_2 \leq a \int_{0}^{y_2} G_{v_0}(v) dv + y_2 =x_2. \end{align*} Hence, the number of $\widehat X^k_0$'s in the interval $[0, x_2]$ is greater than or equal to\break $(a/\varepsilon)\int_0^{y_4} G_{v_1}(v) dv$, with probability greater than $p_2$. Let $\widehat {\bf d}$ be defined as in \eqref{eq:m20.1} but relative to $\widehat I^k$ in place of $I^k$. The two events described in the last two paragraphs hold simultaneously with probability greater than $1-c_1/4$. Then the number of $X^k_0$'s in $[x_1, x_2]$ is greater than or equal to $(a/\varepsilon)\int_{y_3}^{y_4} G_{v_1}(v) dv$. This and \eqref{eq:m21.1} imply that \begin{align}\label{eq:m19.3} \widehat{\bf d}_t([x_1,x_2]) &\geq \frac {\varepsilon\left((a/\varepsilon)\int_{y_5}^{y_6} G_{v_1}(v) dv\right)} {x_2 - x_1 } = \frac {a\int_{y_5}^{y_6} G_{v_1}(v) dv} { \int_0^{y_2} \varphi(z) dz + y_2 - \int_0^{y_1} \varphi(z) dz - y_1 } \nonumber \\ &= \frac {a\int_{y_5}^{y_6} G_{v_1}(v) dv} {y_2 - y_1 + \int_{y_1}^{y_2} \varphi(z) dz } \geq \frac { \int_{y_1}^{y_2} \varphi(z) dz} {y_2 - y_1 + \int_{y_1}^{y_2} \varphi(z) dz } - c_0. \end{align} It follows from \eqref{eq:m21.2} that \begin{align*} \exp ( 2 a (v_0 -y_2) /\sigma^2) = \exp( 2 a v_0 /\sigma^2) - 2ax_2 /\sigma^2 \end{align*} and, therefore, for $v\leq y_2$, \begin{align*} \varphi(v) = \exp ( 2 a (v_0-v) /\sigma^2) -1 \geq \exp ( 2 a (v_0-y_2) /\sigma^2) -1\\ = \exp( 2 a v_0 /\sigma^2)-2ax_2 /\sigma^2 -1. \end{align*} We combine this estimate with \eqref{eq:m19.3} to see that, with probability greater than $1-c_1/4$, \begin{align}\nonumber \widehat {\bf d}_t([x_1,x_2]) &\geq \frac { \int_{y_1}^{y_2} \varphi(v) dv} {y_2 - y_1 + \int_{y_1}^{y_2} \varphi(v) dv } - c_0\\ &\geq \frac { \int_{y_1}^{y_2} (\exp( 2 a v_0 /\sigma^2)-2ax_2 /\sigma^2 -1)dv} {y_2 - y_1 + \int_{y_1}^{y_2} (\exp( 2 a v_0 /\sigma^2)-2ax_2 /\sigma^2 -1)dv } - c_0 \nonumber \\ &= \frac { (y_2-y_1)(\exp( 2 a v_0 /\sigma^2)-2ax_2 /\sigma^2 -1) } {y_2 - y_1 + (y_2-y_1)(\exp( 2 a v_0 /\sigma^2)-2ax_2 /\sigma^2 -1) } - c_0\nonumber \\ &= \frac { \exp( 2 a v_0 /\sigma^2)-2ax_2 /\sigma^2 -1 } { \exp( 2 a v_0 /\sigma^2)-2ax_2 /\sigma^2 } - c_0\nonumber \\ &= \frac { (\sigma^2/2a)(\exp( 2 a v_0 /\sigma^2)-1) -x_2 } { (\sigma^2/2a)(\exp( 2 a v_0 /\sigma^2)-1) -x_2 + \sigma^2/2a} - c_0\nonumber\\ &= \frac { 1 -x_2 } { 1 -x_2 + \sigma^2/2a} - c_0. \label{eq:m21.6} \end{align} The last equality follows from \eqref{eq:m21.4}. Recall that $n_t$ is the number of processes $\widehat Z^k$ alive at time $t$. Note that for any $0\leq t_1< t_2 < \infty$ with $t_2 - t_1 \geq \varepsilon/a$, we have, \begin{align}\label{eq:m19.1} n_{t_2} - n_{t_1} \leq (a/\varepsilon) (t_2-t_1). \end{align} Fix arbitrary $t_1\geq t_0$ and choose $\delta>0$ such that \begin{align}\label{eq:m21.3} \int_0^{v_1} G_{v_1}(v) dv+\delta \leq \int_0^{v_0} G_{v_0}(v) dv. \end{align} We make $\varepsilon_0>0$ smaller, if necessary, so that, by the law of large numbers, if $\varepsilon\in(0,\varepsilon_0)$ then with probability greater than $1-c_1/4$, for all $s_k$ of the form $s_k = k \delta/2$, $k=0, \dots, \lfloor t_1/\delta\rfloor +1$, we have $n_{s_k} < (a/\varepsilon)\left(\int_0^{v_1} G_{v_1}(v) dv+\delta/2\right)$. It follows from \eqref{eq:m19.1} and \eqref{eq:m21.3} that for $\varepsilon< a\delta/2$, \begin{align*} \sup_{0\leq t \leq t_1}n_{t} <(a/\varepsilon)\left( \int_0^{v_1} G_{v_1}(v) dv+\delta\right) \leq (a/\varepsilon) \int_0^{v_0} G_{v_0}(v) dv. \end{align*} Suppose that this event holds. Then, for every $t\leq t_1$, the right edge of the rightmost interval $\widehat I^k_t$ is to the left of \begin{align}\label{eq:m21.5} \varepsilon(a/\varepsilon) \int_0^{v_0} G_{v_0}(v) dv + v_1 = \int_0^{v_0} G_{v_0}(v) dv + v_0 + (v_1 -v_0) = 1 + (v_1- v_0) < 1. \end{align} The second equality in the above formula follows from \eqref{eq:m21.4}. Recall the definitions given before the statement of the theorem. A process $Z^k$ is killed when the right end of the rightmost interval $I^k$ hits 1. Since the processes $\widehat Z^k$ are driven by the same Brownian motions as $Z^k$, \eqref{eq:m21.5} implies that every $Z^k$ has a longer lifetime than $\widehat Z^k$. This implies that ${\bf d}_{t_1}([x_1,x_2]) \geq \widehat {\bf d}_{t_1}([x_1,x_2])$. We combine this with \eqref{eq:m21.6} to conclude that, with probability greater than $1-c_1/2$, \begin{align*} {\bf d}_{t_1}([x_1,x_2]) \geq \widehat {\bf d}_{t_1}([x_1,x_2]) \geq \frac { 1 -x_2 } { 1 -x_2 + \sigma^2/2a} - c_0. \end{align*} A completely analogous argument shows that, with probability greater than $1-c_1/2$, \begin{align*} {\bf d}_{t_1}([x_1,x_2]) \leq \frac { 1 -x_1 } { 1 -x_1 + \sigma^2/2a} + c_0. \end{align*} This completes the proof of the theorem. \end{proof} \section{Discussion} \label{sec:disc} \begin{remark} In the following remarks we will refer to the model analyzed in Section~\ref{sec:pressure} as model (C). We will present another model, which we will call (R). Here, C represents the ``constant'' rate of influx of new particles, and R stands for the ``random'' rate of influx. Model (R) consists of a constant number $n$ of ``particles'' $I^k$ which are confined to the interval $[0,1]$. The $k$-th leftmost ``particle'' is represented by an interval $I^k_t = (X^k_t-\varepsilon/2, X^k_t+\varepsilon/2)$. The intervals $I^k$ and $I^j$ are always disjoint. The processes $X^k$ are driven by independent Brownian motions with the diffusion coefficient $\sigma^2$. When two intervals $I^k$ and $I^j$ collide, they reflect instantaneously. The number of particles $n$ is such that $n\varepsilon=b$, a constant. When $X^k$ hits 1, it jumps to 0. We conjecture that as $\varepsilon \to 0$, the mass density ${\bf d}$ in the stationary regime for this process has the density $(1-x)/(1-x + \sigma^2/(2a))$, just like in model (C), where $a,\sigma$ and $b$ are related by the following formula, \begin{align*} \int_0^1 \frac {1-x} {1-x + \sigma^2/(2a)} = b. \end{align*} Heuristically, we expect processes $X^k$ in model (R) to jump at a more or less constant rate in the stationary regime, so this is why we believe that models (R) and (C) have the same hydrodynamic limit. We chose not to analyze model (R) in this paper as it appears to be harder from the technical point of view while it seems to illustrate the same phenomenon as model (C). \end{remark} \begin{remark} Model (R) is closely related to a model studied by T.~Bodineau, B.~Derrida and J.~Lebowitz (\citealp{BDL}). In their model, one considers a periodic system of $L$ sites with $N$ particles. The particles perform random walks but cannot cross each other---it is the symmetric simple exclusion process. At some fixed edge, the jump rates are no longer symmetric but jumps occur with rates $p$ in one direction and $1-p$ in the other direction. The case $p=1$ corresponds to model (R) described in the previous remark. In the stationary state, the rescaled density varies linearly on the unit line segment, with a discontinuity located where the jump rates are biased. Hence, away from the singularity, the stationary empirical distribution is harmonic for the generator of the single particle process, i.e., Laplacian. This does not apply to the density of mass ${\bf d}$ in our models (C) and (R). \end{remark} \begin{remark} Since the density ${\bf d}$ of the intervals $I^k$ has the form $(1-x)/(1-x + \sigma^2/(2a))$, it is elementary to check that the typical gap size between $I^k$'s is $\varepsilon \sigma^2/(2a(1-x))$. In a model with infinitely small particles $X^k$, the gap size is also $c/(1-x)$ but we do not have any heuristic explanation why the two functions representing the typical gap size should have the same form in both models. \end{remark} \begin{remark}\label{rem:m21.1} We conjecture that the motion of an individual tagged particle $I^k$ in model (C) converges, as $\varepsilon\to 0$, to a deterministic motion with fluctuations having the ``fractional Brownian motion'' structure. In other words, we conjecture that the fluctuations are Gaussian with the local scaling of space and time given by $\Delta x = (\Delta t)^{1/4}$. Our conjecture is inspired by the results in \cite{DGL,H,S,Sw08} on families of one dimensional Brownian motions reflecting from each other. \end{remark} \begin{remark} A $d$-dimensional counterpart of model (C) can be represented as follows. Let $I^k$ be balls with radius $\varepsilon$ and center $X^k$. Our $d$-dimensional model consists of a constant number $n$ of $I^k$'s which are confined to the cube $[0,1]^d$. The balls $I^k$ and $I^j$ are always disjoint. The processes $X^k$ are driven by independent $d$-dimensional Brownian motions with the diffusion coefficient $\sigma^2$. When two balls $I^k$ and $I^j$ collide, they reflect instantaneously. Let $S_\ell$ and $S_r$ be two opposite $(d-1)$-dimensional sides on the boundary of $[0,1]^d$. Balls $I^k$ are pushed into the cube through $S_\ell$ at a constant rate $a$, i.e., $\varepsilon^{-(d-1)}$ balls are pushed into the cube every $\varepsilon/a$ units of time, uniformly over $S_\ell$. Once inside the cube, the balls reflect from all sides except $S_r$. When a ball hits $S_r$, it is removed from the cube. We conjecture that in the stationary regime, when $\varepsilon$ is small, the density of the mass analogous to ${\bf d}$ will be a function of the distance $x$ from $S_\ell$, i.e., a function of depending only on one coordinate. We do not see any obvious reason why the density should have the form $(1-x)/(1-x + \sigma^2/(2a))$. In relation to Remark \ref{rem:m21.1}, we conjecture that the motion of a tagged particle in the present model is diffusive, with the diffusion coefficient depending on $x$. If this is true, it means that the ``pressure'' applied to particles in one direction can have a dampening effect on the size of oscillations of an individual particle in orthogonal directions. \end{remark} \bibliographystyle{alea2}
\section{Introduction} In a recent paper \cite{IAH} a proposal has been made to study nucleon--nucleon (NN) potentials from a first principle QCD approach. In this field theoretic framework, potentials are obtained through the Schr\"odinger operator applied to Bethe--Salpeter (BS) wave functions defined by \begin{eqnarray} \varphi_E(\vec r) &=& \langle 0 \vert N(\vec x+\vec r, t)N(\vec x , t) \vert 2{\rm N}, E\rangle\,, \end{eqnarray} where $\vert 2{\rm N}, E\rangle$ is a QCD eigenstate with energy $E$ (suppressing here other quantum numbers), and $N$ is a nucleon interpolating operator made of 3 quarks such as $N(x) =\epsilon^{abc}q^a(x)q^b(x)q^c(x)$. Such wave functions have been measured through numerical simulations of the lattice regularized theory \cite{IAH,AHI1,AHI2,IAH2}. Although many conceptual questions remain to be resolved, the corresponding potentials indeed qualitatively resemble phenomenological NN potentials which are widely used in nuclear physics. The force at medium to long distance ($r\ge 2$ fm) is shown to be attractive. This feature has long well been understood in terms of pion and other heavier meson exchanges. At short distance, a characteristic repulsive core is reproduced by the lattice QCD simulation \cite{IAH}. No simple theoretical explanation, however, exists so far for the origin of the repulsive core. For an approach based on string theories, see ref. \cite{string}. By writing \begin{eqnarray} \langle 0 \vert N(\vec x+\vec r, t) N(\vec x , t) &=& \sum_{n=0}^\infty \int \frac{\mathrm{d} E}{2E} \langle 2{\rm N}, n\pi, E\vert f_n(\vec r, E)\,, \end{eqnarray} where $\vert 2{\rm N}, n\pi, E\rangle$ is a state with the energy $E$ containing two nucleons and $n$ pions (and/or nucleon-antinucleon pairs), we see that $\varphi_E(\vec r) = f_0(\vec r, E)$. (Our normalization is $\langle 2{\rm N}, n\pi, E \vert 2{\rm N}, n'\pi, E'\rangle =2E\delta_{nn'}\delta(E-E')$.) We may thus interpret the wave function $\varphi_E(\vec r)$ as an amplitude to find the QCD eigenstate $\vert 2{\rm N}, E\rangle$ in $N(\vec x+\vec r, t) N(\vec x , t) \vert 0 \rangle$. The behavior of the wave functions $\varphi_E(\vec r)$ at short distances ($r=\vert \vec r\vert$) are encoded in the operator product expansion (OPE) of $ N(\vec x+\vec r, t) N(\vec x , t)$. An OPE analysis \cite{ABW1} of BS wave functions in the case of a toy model, the Ising field theory in 2--dimensions, successfully described the analytically known behavior. In this case the limiting short distance behavior of the potential does not depend on the energy (rapidity) of the state, and further it only mildly depends on energy (for low energies) at distances of the order of the Compton wave length of the particles. In this report we perform an operator product expansion (OPE) analysis of NN BS wave functions in QCD, with the aim to theoretically better understand the repulsive core of the NN potential, (at least that of the measured BS potential). Thanks to the property of asymptotic freedom of QCD the form of leading short distance behavior of the coefficent functions can be computed using perturbation theory. A short summary of our results has been published in ref. \cite{ABW2}\footnote{Unfortunately the results for $\beta^{01}$ and $\beta^{10}$ as given in ref. \cite{ABW2} differ (incorrectly) from (\ref{summary2}) by a factor of 2.}. In sect.~2 we start with some general considerations on BS potentials, and sect.~3 presents some standard renormalization group considerations. The anomalous dimensions of 3-- and 6--quark operators are computed in sect.~4. Finally in sects.~5,~6 we discuss the application of the results to NN potentials. In appendix~C we make a similar analysis for the $I=2$ two pion system. For the convenience of the reader we give a brief summary of our results here. The OPE analysis shows that the NN central potential at short distance behaves as \begin{eqnarray} V_c^{SI} (r) &\simeq& C_E \frac{(-\log r )^{\beta^{SI} -1}}{r^2} \label{summary1} \end{eqnarray} for the S--state ($L=0$) , where $S$ and $I$ are total spin and isospin of the NN system, respectively, and $\beta^{SI}$ is negative and explicitly obtained as \begin{eqnarray} \beta^{01} &=& - \frac{6}{33-2N_\mathrm{f}}, \quad \beta^{10} = - \frac{2}{33-2N_\mathrm{f}}\,, \label{summary2} \end{eqnarray} where $N_\mathrm{f}$ is the number of quark flavors, and the overall coefficient $C_E$ depends on the energy $E$. Unfortunately the OPE analysis is not as conclusive as that in the toy model referred to above, in particular the sign of $C_E$ is not determined by perturbative cosiderations alone. The latter requires additional non-perturbative knowledge of matrix elements of composite operators. A crude estimation using the non-relativistic quark model indicates that $C_E$ is positive, which implies a repulsive core with a potential diverging a little weaker than the generically expected $r^{-2}$ at short distances. \section{Operator Product Expansion and potentials at short distance in 3 dimensions} \label{sec:3dim} In this section we discuss the application of the operator product expansion (OPE) to the determination of the short distance behavior of the BS potential. We consider the equal time Bethe--Salpeter (BS) wave function defined by \begin{eqnarray} \varphi_{AB}^E(\vec r) &=& \langle 0 \vert O_A(\vec r/2, 0) O_B(-\vec r/2, 0) \vert E \rangle\,, \end{eqnarray} where $\vert E \rangle$ is an eigen-state of the system with energy $E$, and $O_A$, $O_B$ are some operators of the system. Here we suppress other quantum numbers of the state $\vert E\rangle $ for simplicity. Using the OPE of $O_A$ and $O_B$ \begin{eqnarray} O_A (\vec r/2, 0) O_B(-\vec r/2,0) &\simeq& \sum_C D_{AB}^C(\vec r) O_C(\vec 0, 0)\,, \label{ope22} \end{eqnarray} we have \begin{eqnarray} \varphi_{AB}^E(\vec r) &\simeq& \sum_C D_{AB}^C(\vec r) \langle 0 \vert O_C(\vec 0, 0)\vert E \rangle\,. \end{eqnarray} Note that $\vec r$ dependence appears solely in $D_{AB}^C(\vec r)$ while the $E$ dependence exists only in $\langle 0 \vert O_C(\vec 0,0) \vert E \rangle$. As we will see, in the $r=\vert \vec r\vert \rightarrow 0$ limit, the coefficient function behaves as \begin{eqnarray} D_{AB}^C (\vec r) \simeq r^{\alpha_C} (-\log r)^{\beta_C} f_C(\theta,\phi)\,, \label{alphabeta} \end{eqnarray} where $\theta,\phi$ are angles in the polar coordinates of $\vec r$, so that \begin{eqnarray} \varphi_{AB}^E (\vec r) \simeq \sum_C r^{\alpha_C}(-\log r)^{\beta_C} f_C(\theta,\phi)D_C(E)\,, \quad D_C(E) = \langle 0 \vert O_C (\vec 0,0)\vert E \rangle\,. \end{eqnarray} We now assume that $C$ has the largest contribution at small $r$: \begin{eqnarray} \alpha_C &<& \alpha_{C'} \quad \mbox{or} \\ \alpha_C &=& \alpha_{C'}, \quad \beta_C > \beta_{C'}\,. \end{eqnarray} for $^\forall C'\not= C$. The potential can be calculated from this wave function. As will be seen later, $\alpha_C=\alpha_{C^\prime} =0$ for the NN case in QCD. Furthermore states with zero orbital angular momentum ($L=0$) dominates in the OPE, so that the wave function at short distance is given by \begin{eqnarray} \varphi_{AB}^E(r) \simeq \left[ (-\log r)^{\beta_C} D_C(E) + (-\log r)^{\beta_{C^\prime}} D_{C^\prime}(E) \right] \end{eqnarray} with $\beta_C > \beta_{C^\prime}$. Using \begin{eqnarray} \nabla^2 (-\log r)^{\beta} &=& -\beta (-\log r)^{\beta-1} \left[ 1- \frac{\beta-1}{-\log r}\right] r^{-2} \,, \end{eqnarray} we obtain the following classification of the short distance behavior of the potential. \begin{enumerate} \item $\beta_C \not=0$: The potential at short distance is energy independent and becomes \begin{eqnarray} V(r) &\simeq& -\frac{\beta_C}{r^2(-\log r)}\,, \end{eqnarray} which is attractive for $\beta_C >0 $ and repulsive for $\beta_C < 0$. \item $\beta_C =0$: In this case we have \begin{eqnarray} V(r) &\simeq& \frac{D_{C'}(E)}{D_C(E)} \left(\frac{-\beta_{C'}}{r^2}\right) (-\log r)^{\beta_{C^\prime}-1}\,. \end{eqnarray} The sign of the potential at short distance depends on $ -\beta_{C^\prime}D_{C'}(E) / D_C(E)$. \end{enumerate} If there are two or more operators which have the largest contribution at short distance, we have \begin{eqnarray} \varphi_{AB}^E(x) &=& (-\log r)^{\beta_C}( D_{C_1}(E) + D_{C_2}(E)+\cdots)\,. \end{eqnarray} In this case, the above analysis can be applied just by replacing $D_C(E) \rightarrow D_{C_1}(E) + D_{C_2}(E)+\cdots$. On the lattice, we do not expect divergences at $r=0$ due to lattice artifacts at short distance. The above classification holds at $a \ll r \ll 1\,{\rm fm}$, while the potential becomes finite even at $r=0$ on the lattice. \section{Renormalization group analysis and operator product expansion} \subsection{Renormalization group equation for composite operators} In QCD, using dimensional regularization in $D=4-2\epsilon$ dimensions, bare local composite operators $O^{(0)}_A(x)$ are renormalized \cite{RG} according to\footnote{We note that we are considering the massless theory here since quark masses play no role in our analysis.} \begin{equation} O^{({\rm ren})}_A(x)=Z_{AB}(g,\epsilon)\,O^{(0)}_B(x). \end{equation} (Summation of repeated indices is assumed throughout this paper.) The meaning of the above formula is that we obtain finite results if we insert the right hand side into any correlation function, provided we also renormalize the bare QCD coupling $g_0$ and the quark and gluon fields. For example, in the case of an $n$--quark correlation function with operator insertion, which we denote by ${\cal G}^{(0)}_{n;A}(g_0,\epsilon)$ (suppressing the dependence on the quark momenta and other quantum numbers) we have \begin{equation} {\cal G}^{({\rm ren})}_{n;A}(g,\mu)=Z_{AB}(g,\epsilon)\, Z_F^{-n/2}(g,\epsilon)\, {\cal G}^{(0)}_{n;B}(g_0,\epsilon). \end{equation} We recall from renormalization theory that for the analogous $n$--quark correlation function (without any operator insertion) we have \begin{equation} {\cal G}^{({\rm ren})}_n(g,\mu)=Z_F^{-n/2}(g,\epsilon)\, {\cal G}^{(0)}_n(g_0,\epsilon), \end{equation} where the coupling renormalization is given by \begin{equation} g_0^2=\mu^{2\epsilon}\,Z_1(g,\epsilon)\,g^2. \end{equation} The renormalization constant $Z_1$ in the minimal subtraction (MS) scheme we are using has pure pole terms only: \begin{equation} Z_1(g,\epsilon)=1-\frac{\beta_0g^2}{\epsilon}-\frac{\beta_1g^4}{2\epsilon} +\frac{\beta_0^2g^4}{\epsilon^2}+\mathrm{O}(g^6), \end{equation} where \begin{equation} \beta_0=\frac{1}{16\pi^2}\left\{\frac{11}{3}N-\frac{2}{3}N_f\right\}, \qquad\quad \beta_1=\frac{1}{256\pi^4}\left\{\frac{34}{3}N^2-\left( \frac{13}{3}N-\frac{1}{N}\right)N_f\right\}. \end{equation} Similarly for the fermion field renormalization constant, we have \begin{equation} Z_F(g,\epsilon)=1-\frac{\gamma_{F0}g^2}{2\epsilon}+\mathrm{O}(g^4), \end{equation} where $\gamma_{F0}$ is given by (\ref{wfrc}). The gluon field renormalization constant is also similar, but we do not need it here. Finally the matrix of operator renormalization constants is of the form \begin{equation} Z_{AB}(g,\epsilon)=\delta_{AB}-\frac{\gamma^{(1)}_{AB}g^2}{2\epsilon} +\mathrm{O}(g^4). \end{equation} The renormalization group (RG) expresses the simple fact that bare quantities are independent of the renormalization scale $\mu$. Introducing the RG differential operator \begin{equation} {\cal D}=\mu\frac{\partial}{\partial\mu}+\beta(g)\,\frac{\partial}{\partial g} \end{equation} the RG equation for $n$--quark correlation functions can be written as \begin{equation} \left\{{\cal D}+\frac{n}{2}\gamma_F(g)\right\}\, {\cal G}^{({\rm ren})}_n(g,\mu)=0, \end{equation} where the RG beta function is \begin{equation} \beta(g)=\epsilon g +\beta_D(g,\epsilon) =\epsilon g-\frac{\epsilon g}{1+\frac{g}{2}\, \frac{\partial\ln Z_1}{\partial g}}=-\beta_0g^3-\beta_1g^5+\mathrm{O}(g^7), \end{equation} where $\beta_D(g,\epsilon)$ is the beta function in $D$ dimenions and the RG gamma function (for quark fields) is \begin{equation} \gamma_F(g)=\beta_D(g,\epsilon)\,\frac{\partial\ln Z_F} {\partial g}=\gamma_{F0}\,g^2+\mathrm{O}(g^4). \end{equation} It is useful to introduce the RG invariant lambda-parameter $\Lambda$ by taking the Ansatz \begin{equation} \Lambda=\mu\,{\rm e}^{f(g)} \end{equation} and requiring ${\cal D}\Lambda=0$. The solution is the lambda-parameter in the MS scheme ($\Lambda_{\rm MS}$) if the arbitrary integration constant is fixed by requiring that for small coupling \begin{equation} f(g)=-\frac{1}{2\beta_0g^2}-\frac{\beta_1}{2\beta_0^2}\,\ln(\beta_0g^2) +\mathrm{O}(g^2). \end{equation} Finally the RG equations for $n$--quark correlation functions with operator insertion are of the form \begin{equation} \left\{{\cal D}+\frac{n}{2}\gamma_F(g)\right\}\, {\cal G}^{({\rm ren})}_{n;A}(g,\mu)-\gamma_{AB}(g) {\cal G}^{({\rm ren})}_{n;B}(g,\mu)=0, \end{equation} where \begin{equation} \gamma_{AB}(g)=-Z_{AC}\beta_D(g,\epsilon)\frac {\partial Z^{-1}_{CB}}{\partial g}=\gamma^{(1)}_{AB}g^2+\mathrm{O}(g^4). \end{equation} \subsection{OPE and RG equations} Let us recall the operator product expansion (\ref{ope22}) \begin{equation} O_1(y/2)O_2(-y/2)\simeq D_B(y)\,O_B(0). \label{ope33} \end{equation} We will need it in the special case where the operators $O_1,O_2$ on the left hand side are nucleon operators and the set of operators $O_B$ on the right hand side are local 6--quark operators of engineering dimension 9 and higher. All operators in (\ref{ope33}) are renormalized ones, but from now on we suppress the labels $^{({\rm ren})}$. As we will see, the nucleon operators are renormalized diagonally as \begin{equation} O_1=\zeta_1(g,\epsilon)\,O^{(0)}_1,\qquad\qquad O_2=\zeta_2(g,\epsilon)\,O^{(0)}_2, \end{equation} and we can define the corresponding RG gamma functions by \begin{equation} \gamma_{1,2}(g)=\beta_D(g,\epsilon)\, \frac{\partial\ln\zeta_{1,2}}{\partial g}=\gamma_{1,2}^{(1)}g^2+\mathrm{O}(g^4). \end{equation} Next we write down the bare version of (\ref{ope33}) (in terms of bare operators and bare coefficient functions): \begin{equation} O^{(0)}_1(y/2)O_2^{(0)}(-y/2)\simeq D^{(0)}_B(y)\,O^{(0)}_B(0). \label{ope33bare} \end{equation} Comparing (\ref{ope33}) to (\ref{ope33bare}), we can read off the renormalization of the coefficient functions: \begin{equation} D_B(y)=\zeta_1(g,\epsilon)\zeta_2(g,\epsilon)D^{(0)}_A(y)\, Z^{-1}_{AB}(g,\epsilon) \end{equation} and using the $\mu$-independence of the bare coefficient functions we can derive the RG equations satisfied by the renormalized ones: \begin{equation} {\cal D}D_B(g,\mu,y)+D_A(g,\mu,y)\,\tilde\gamma_{AB}(g)=0, \label{RG33} \end{equation} where the effective gamma function matrix is defined as \begin{equation} \tilde\gamma_{AB}(g)=\gamma_{AB}(g)-\left[\gamma_1(g)+\gamma_2(g)\right] \,\delta_{AB}. \end{equation} \subsection{Perturbative solution of the RG equation and factorization of OPE} We want to solve the vector partial differential equation (\ref{RG33}) and for this purpose it is useful to introduce $\hat U_{AB}(g)$, the solution of the matrix ordinary differential equation \begin{equation} \beta(g)\,\frac{{\rm d}}{{\rm d}g}\,\hat U_{AB}(g)=\tilde\gamma_{AC}(g) \,\hat U_{CB}(g) \label{hatU} \end{equation} and its matrix inverse $U_{AB}(g)$. We will assume that the coefficient functions are dimensionless and have the perturbative expansion \begin{equation} D_A(g,\mu,y)=D_A(g;\mu r)=D_{A;0}+g^2D_{A;1}(\mu r)+\mathrm{O}(g^4), \label{Dpert} \end{equation} where $r=\vert y\vert$. For the case of operators with higher engineering dimension $9+\alpha$ the coefficients are of the form $r^\alpha$ times dimensionless functions and the analysis is completely analogous and can be done independently, since in the massless theory operators of different dimension do not mix. (In the full theory quark mass terms are also present, but they correspond to higher powers in $r$ and therefore can be neglected.) We will also assume that the basis of operators has been chosen such that the 1-loop mixing matrix is diagonal: \begin{equation} \tilde\gamma_{AB}(g)=2\beta_0\,\beta_A\,g^2\,\delta_{AB}+\mathrm{O}(g^4). \end{equation} In such a basis the solution of (\ref{hatU}) in perturbation theory takes the form \begin{equation} \hat U_{AB}(g)=\left\{\delta_{AB}+R_{AB}(g)\right\}\,g^{-2\beta_B}, \label{hatUpert} \end{equation} where $R_{AB}(g)=\mathrm{O}(g^2)$, with possible multiplicative $\log g^2$ factors, depending on the details of the spectrum of 1-loop eigenvalues $\beta_A$. Having solved (\ref{hatU}) we can write down the most general solution of (\ref{RG33}): \begin{equation} D_B(g;\mu r)=F_A(\Lambda r)\,U_{AB}(g). \end{equation} Here the vector $F_A$ is RG-invariant. Introducing the running coupling $\bar g$ as the solution of the equation \begin{equation} f(\bar g)=f(g)+\ln(\mu r)=\ln(\Lambda r) \end{equation} $F_B$ can be rewritten as \begin{equation} F_B(\Lambda r)=D_A(\bar g;1)\,\hat U_{AB}(\bar g). \end{equation} Since, because of asymptotic freedom (AF), for $r\to0$ also $\bar g\to0$ as \begin{equation} \bar g^2\approx-\frac{1}{2\beta_0\ln(\Lambda r)}, \end{equation} $F_B$ can be calculated perturbatively using (\ref{Dpert}) and (\ref{hatUpert}). Putting everything together, we find that the right hand side of the operator product expansion (\ref{ope33}) can be rewritten: \begin{equation} O_1(y/2)O_2(-y/2)\simeq F_B(\Lambda r)\,\tilde O_B(0), \label{ope34} \end{equation} where \begin{equation} \tilde O_B=U_{BC}(g)\,O_C. \end{equation} There is a factorization of the operator product into perturbative and non-perturbative quantities: $F_B(\Lambda r)$ is perturbative and calculable (for $r\to0$) thanks to AF, whereas the matrix elements of $\tilde O_B$ are non-perturbative (but $r$-independent). An operator $O_B$ first occurring at $\ell_B$-loop order on the right hand side of (\ref{ope33}) and corresponding to normalized 1-loop eigenvalue $\beta_B$ has coefficient $F_B(\Lambda r)$ with leading short distance behavior \begin{equation} F_B(\Lambda r)\approx \bar g^{2(\ell_B-\beta_B)}\approx \left(-2\beta_0\ln(\Lambda r)\right)^{\beta_B-\ell_B}. \end{equation} In principle, an operator with very large $\beta_B$, even if it is not present in the expansion at tree level yet, might be important at short distances. This is why it is necessary to calculate the full 1-loop spectrum of $\beta_B$ eigenvalues. As we shall see, no such operators exist in our cases, and therefore operators with non-vanishing tree level coefficients are dominating at short distances. The corresponding coefficient functions have leading short distance behavior given by \begin{equation} F_B(\Lambda r)\approx D_{B;0}\,\left(-2\beta_0\ln(\Lambda r)\right)^{\beta_B}. \end{equation} A similar analysis in the case of operators of dimension $9+\alpha$ leads to the result (\ref{alphabeta}). \section{OPE and Anomalous dimensions for two nucleons} \subsection{OPE of two nucleon operators at tree level} The general form of a gauge invariant local 3--quark operator is given by \begin{eqnarray} B^F_\Gamma (x) \equiv B^{fgh}_{\alpha\beta\gamma}(x) = \varepsilon^{abc} q^{a,f}_\alpha(x) q^{b,g}_\beta(x) q^{c,h}_\gamma (x)\,, \label{baryonop} \end{eqnarray} where $\alpha,\beta,\gamma$ are spinor, $f,g,h$ are flavor, $a,b,c$ are color indices of the (renormalized) quark field $q$. The color index runs from 1 to $N=3$, the spinor index from 1 to 4, and the flavor index from 1 to $N_\mathrm{f}$. In this paper a summation over a repeated index is assumed, unless otherwise stated. Note that $B^{fgh}_{\alpha\beta\gamma}$ is symmetric under any interchange of pairs of indices (e.g. $B^{fgh}_{\alpha\beta\gamma}=B^{gfh}_{\beta\alpha\gamma}$) because the quark fields anticommute. For simplicity we sometimes use the notation such as $F=fgh$ and $\Gamma=\alpha\beta\gamma$ as indicated in (\ref{baryonop}). The nucleon operator is constructed from the above operators as \begin{eqnarray} B^f_\alpha(x) = \left(P_{+4}\right)_{\alpha\alpha'} B_{\alpha'\beta\gamma}^{fgh} (C\gamma_5)_{\beta\gamma}(i\tau_2)^{gh}\,, \end{eqnarray} where $P_{+4} = (1+\gamma_4)/2$ is the projection to the large spinor component, $C=\gamma_2\gamma_4$ is the charge conjugation matrix, and $\tau_2$ is the Pauli matrix in the flavor space (for $N_\mathrm{f}=2$) given by $(i\tau_2)^{fg} = \varepsilon^{fg}$. Both $C\gamma_5$ and $i\tau_2$ are anti-symmetric under the interchange of two indices, so that the nucleon operator has spin $1/2$ and isospin $1/2$. Although the explicit form of the $\gamma$ matrices is unnecessary in principle, we find it convenient to use a (chiral) convention given by \begin{eqnarray} \gamma_k &=& \left( \begin{array}{cc} 0 & i\sigma_k \\ -i\sigma_k & 0 \end{array} \right)\,, \ \gamma_4 = \left( \begin{array}{cc} 0 & {\bf 1} \\ {\bf 1} & 0 \\ \end{array} \right)\,, \ \gamma_5 =\gamma_1\gamma_2\gamma_3\gamma_4 = \left( \begin{array}{cc} {\bf 1}& 0 \\ 0 & -{\bf 1} \\ \end{array} \right)\,. \end{eqnarray} As discussed in the previous section, the OPE at the tree level (generically) dominates at short distance. The OPE of two nucleon operators given above at tree level becomes \begin{eqnarray} B^f_{\alpha}(x+y/2) B^g_{\beta}(x-y/2) &=& B^f_{\alpha}(x) B^g_{\beta}(x) +\frac{y^\mu}{2}\left\{ \partial_\mu[ B^f_\alpha (x)] B^g_\beta(x) - B^f_\alpha (x)\partial_\mu[ B^g_\beta(x)] \right\}\nonumber \\ &+&\frac{y^\mu y^\nu}{8} \left\{\partial_\mu\partial_\nu [B^f_\alpha(x) B^g_\beta(x) ] -4\partial_\mu B^f_\alpha(x)\partial_\nu B^g_\beta(x) \right\} +\cdots . \end{eqnarray} For the two-nucleon operator with either the combination $[\alpha\beta]$, $\{fg\}$ ($S=0$) or the combination $\{\alpha\beta\}$, $[fg]$ ($S=1$), terms odd in $y$ vanish in the above OPE, so that only even $L$ contributions appear. These 6--quark operators are anti-symmetric under the exchange $(\alpha,f)\leftrightarrow (\beta,g)$. On the other hand, for two other operators with $( [\alpha\beta], [fg])$ or $(\{\alpha\beta\},\{fg\})$, which are symmetric under the exchange, terms even in $y$ vanish in the OPE and only odd $L$'s contribute. Knowing the anomalous dimensions of the 6--quark operators appearing in the OPE, which will be calculated later in this section, the OPE at short distance ($r=\vert \vec y\vert \ll 1$, $y_4=0$) becomes \begin{eqnarray} B^f_{\alpha}(x+y/2) B^g_{\beta}(x-y/2) &\simeq& \sum_A c_A(r) O^{fg,A}_{\alpha\beta}(x) +\sum_B d_B(r) y^k y^l O^{fg,B}_{\alpha\beta,kl} (x) \nonumber \\ &+&\sum_C e_C(r) y^k O^{fg,C}_{\alpha\beta,k}(x) +\cdots \,, \end{eqnarray} where the coefficient functions behave as \begin{eqnarray} c_A (r) &\simeq& (-\log r)^{\beta_A}\,, \quad d_B (r) \simeq (-\log r)^{\beta_B}\,, \quad e_C (r) \simeq (-\log r)^{\beta_C}\,, \end{eqnarray} and $\beta_{A,B,C}$ are related to the anomalous dimensions of the 6--quark operators $O^{fg,A}_{\alpha\beta}$, of those with two derivatives $O^{fg,B}_{\alpha\beta,kl}$ and of those with one derivative $O^{fg,C}_{\alpha\beta,k}$. The wave function defined through the eigenstate $\vert E\rangle$ is given by \begin{eqnarray} \varphi_E^{\rm even}(y) &=& \langle 0 \vert B^f_{\alpha}(x+y/2) B^g_{\beta}(x-y/2)\vert E\rangle \nonumber\\ &\simeq& \sum_A c_A (r) \langle 0 \vert O^{fg,A}_{\alpha\beta}(x) \vert E\rangle + \sum_B d_B (r) y^k y^l \langle 0 \vert O^{fg,B}_{\alpha\beta,kl} (x) \vert E\rangle+\mathrm{O}(y^4)\,, \end{eqnarray} for the anti-symmetric states, while \begin{eqnarray} \varphi_E^{\rm odd}(y) &=& \langle 0 \vert B^f_{\alpha}(x+y/2) B^g_{\beta}(x-y/2)\vert E\rangle \simeq \sum_C e_C (r) y^k \langle 0 \vert O^{fg,C}_{\alpha\beta,k} (x) \vert E\rangle+\mathrm{O}(y^3) \end{eqnarray} for the symmetric states. In this paper, we consider only 6--quark operators without derivatives and calculate the corresponding anomalous dimensions. \subsection{General formula for the divergent part at 1-loop} Following the previous section, we define the renormalization factor $Z_X$ of a $k$--quark operator $X=[q^k]$ through the relation \begin{eqnarray} [q^k]^{\mathrm{ren}} &=& Z_X [q_0^k] = Z_X Z_F^{k/2}[q^k]\,, \end{eqnarray} where $q_0$($q$) is the bare (renormalized) quark field. The wave function renormalization factor for the quark field is given at 1-loop by \begin{eqnarray} Z_F &=& 1 + g^2 Z_F^{(1)} \,, \quad Z_F^{(1)} = -\frac{\lambda C_F }{16\pi^2\epsilon} \label{wfrc} \end{eqnarray} where $\lambda$ is the gauge parameter and $C_F =\frac{N^2-1}{2N}$. At 1-loop the renormalization of simple $k$--quark operators (those involving no gauge fields) is given by the divergent parts of diagrams involving exchange of a gluon between any pair of quark fields. The 1-loop correction to the insertion of an operator $q^{a,f}_\alpha(x) q^{b,g}_\beta (x)$ in any correlation function involving external quarks is expressed as the contraction of \begin{eqnarray} q^{a,f}_\alpha(x) q^{b,g}_\beta(x) \frac{1}{2!} \int \mathrm{d}^D y\, \mathrm{d}^D z \, A_\mu^A(y) A_\nu^B(z) [\bar q^{f_1}(y) ig T^A \gamma_\mu q^{f_1}(y)] [\bar q^{g_1}(z) ig T^B \gamma_\nu q^{g_1}(z)] \label{eq:op_1loop} \end{eqnarray} where $\mathrm{tr}\, T^A T^B =\delta_{AB}/2$ in our normalization. Since two identical contributions cancel the $2!$ in the denominator, the contraction at 1-loop is given by \begin{eqnarray} -g^2 (T^A)_{aa_1}(T^A)_{bb_1}\int \mathrm{d}^Dy\, \mathrm{d}^D z\, && \left[S_F(x-y)\gamma_\mu q(y)\right]_{\alpha}^{a_1f_1} G_{\mu\nu}(y-z)\nonumber \\ &\times& \left[S_F(x-z)\gamma_\nu q(z)\right]_{\beta}^{b_1g_1} \end{eqnarray} where the free quark and gauge propagators are given in momentum space as \begin{eqnarray} S_F(p) &=& \frac{ - i p\kern-1ex / + m}{p^2+ m^2}\,, \quad G_{\mu\nu}(k) = \frac{1}{k^2} \left[ g_{\mu\nu} - (1-\lambda)\frac{k_\mu k_\nu}{k^2}\right]\,. \end{eqnarray} The above contribution can be written as \begin{eqnarray} \frac{g^2}{2N}\{\delta_{aa_1}\delta_{bb_1}-N\delta_{ab_1}\delta_{a_1b}\} \int \frac{\mathrm{d}^D p\, \mathrm{d}^D q}{(2\pi)^{2D}} T_{\alpha\alpha_1,\beta\beta_1}(p,q)\, q^{a_1f_1}_{\alpha_1}(p)\mathrm{e}^{ipx} q^{b_1g_1}_{\beta_1}(q)\mathrm{e}^{iqx} \label{eq:contraction} \end{eqnarray} where \begin{eqnarray} T_{\alpha\alpha_1,\beta\beta_1}(p,q) &=& \int \frac{\mathrm{d}^D k}{(2\pi)^D} \left[S_F(p+k) \gamma_\mu\right]_{\alpha\alpha_1} G_{\mu\nu}(k) \left[S_F(q-k) \gamma_\nu\right]_{\beta\beta_1}\,, \end{eqnarray} whose divergent part is independent of the momenta $p,q$ and is given by \begin{eqnarray} T_{\alpha\alpha_1,\beta\beta_1}(0,0) &=& \frac{1}{16\pi^2}\frac{1}{\epsilon} \left[ -\frac{1}{4} \sum_{\mu\nu} \sigma_{\mu\nu} \otimes \sigma_{\mu\nu} + \lambda 1\otimes1\right]_{\alpha\alpha_1,\beta\beta_1} \end{eqnarray} with $\sigma_{\mu\nu} =\frac{i}{2}\left[\gamma_\mu,\gamma_\nu\right]$. We then obtain the divergent part of the 1-loop contribution as \begin{eqnarray} \left[ q^{a,f}_\alpha(x) q^{b,g}_\beta (x)\right]^{\rm 1-loop, div} &=& \frac{g^2}{32 N\pi^2}\frac{1}{\epsilon} \left[ ({\bf T}_0 + \lambda {\bf T}_1) \cdot q^a(x)\otimes q^b (x)\right]_{\alpha,\beta}^{fg} \label{eq:1-loopT} \end{eqnarray} where (bold--faced symbols represent matrices in flavor and spinor space) \begin{eqnarray} ({\bf T}_0)^{f f_1,g g_1}_{\alpha\alpha_1,\beta\beta_1} &=& -\frac{1}{4}\sum_{\mu\nu}\left\{ \boldsymbol{\sigma}_{\mu\nu}\otimes\boldsymbol{\sigma}_{\mu\nu} + N \boldsymbol{\sigma}_{\mu\nu}\tilde\otimes\boldsymbol{\sigma}_{\mu\nu} \right\}^{f f_1,g g_1}_{\alpha\alpha_1,\beta\beta_1}\,, \\ ({\bf T}_1)^{f f_1,g g_1}_{\alpha\alpha_1,\beta\beta_1} &=& \left\{{\bf 1}\otimes{\bf 1} + N{\bf 1}\tilde\otimes{\bf 1}\right\}^{f f_1,g g_1}_{\alpha\alpha_1,\beta\beta_1}\,. \label{eq:T1} \end{eqnarray} Here we use the notation \begin{eqnarray} \{{\bf X}\otimes{\bf Y}\}^{f f_1,g g_1}_{\alpha\alpha_1,\beta\beta_1} &=& {\bf X}^{f f_1}_{\alpha\alpha_1}{\bf Y}^{g g_1}_{\beta\beta_1} \qquad \{{\bf X}\tilde\otimes{\bf Y}\}^{f f_1,g g_1}_{\alpha\alpha_1,\beta\beta_1} = {\bf X}^{g f_1}_{\beta\alpha_1} {\bf Y}^{f g_1}_{\alpha\beta_1}\,, \\ \{\boldsymbol{\sigma}_{\mu\nu}\}^{f g}_{\alpha\beta}&=&\delta^{fg} (\sigma_{\mu\nu})_{\alpha\beta}, \quad \{{\bf 1}\}^{f g}_{\alpha\beta} = \delta^{fg}\delta_{\alpha\beta}\,. \end{eqnarray} Using the following Fierz identities for spinor indices \begin{eqnarray} -\frac{1}{4}\sum_{\mu\nu} \sigma_{\mu\nu}\otimes\sigma_{\mu\nu} &=& P_R\otimes P_R + P_L\otimes P_L - 2( P_R\tilde\otimes P_R + P_L\tilde\otimes P_L)\,, \label{eq:fierz1}\\ -\frac{1}{4}\sum_{\mu\nu} \sigma_{\mu\nu}\tilde\otimes \sigma_{\mu\nu} &=& P_R\tilde\otimes P_R + P_L\tilde\otimes P_L - 2( P_R\otimes P_R + P_L\otimes P_L)\,, \label{eq:fierz2} \end{eqnarray} where $P_R,P_L$ are the chiral projectors \begin{eqnarray} P_R&=&\frac12(1+\gamma_5)\,,\,\,\,\,\, P_L =\frac12(1-\gamma_5)\,, \end{eqnarray} we can simplify ${\bf T}_0$ as \begin{eqnarray} ({\bf T}_0)^{f f_1,g g_1}_{\alpha\alpha_1,\beta\beta_1} &=& \delta^{ff_1}\delta^{gg_1} \left[ \delta_{\alpha\alpha_1}\delta_{\beta\beta_1} - 2\delta_{\beta\alpha_1}\delta_{\alpha\beta_1}\right] +N\delta^{gf_1}\delta^{fg_1} \left[ \delta_{\beta\alpha_1}\delta_{\alpha\beta_1} -2\delta_{\alpha\alpha_1}\delta_{\beta\beta_1} \right]\nonumber \\ \label{eq:T0} \end{eqnarray} where either $\alpha_1,\beta_1 \in \{1,2\}$(right-handed) or $\alpha_1,\beta_1 \in \{3,4\}$(left-handed) due to the chiral projections in eqs. (\ref{eq:fierz1}) and (\ref{eq:fierz2}). In our following calculation of the 1-loop anomalous dimensions, eq.~(\ref{eq:1-loopT}) together with eqs.~(\ref{eq:T0}) and (\ref{eq:T1}) are the key equations. \subsection{Renormalization of local 3--quark operators at 1-loop} In this subsection we calculate the anomalous dimensions of general 3--quark operators at 1-loop. In terms of the renormalization factor defined as \begin{eqnarray} B_3^{\rm renor.} &=& \zeta [q_0^3] = \zeta Z_F^{3/2}[q^3] , \quad \zeta = 1 + g^2 (\zeta^{(1)} +\zeta_\lambda^{(1)} )+\dots\,, \end{eqnarray} where $\zeta^{(1)}$ ( $\zeta_\lambda^{(1)}$ ) is the $\lambda$--independent (dependent) part at 1-loop, the divergent part of the insertion of the 3--quark operator $B^F_\Gamma=B_{\alpha\beta\gamma}^{fgh}$ defined in (\ref{baryonop}) at 1-loop is given by a linear combination of insertion of baryon operators, and (with a slight abuse of notation) we express this as \begin{eqnarray} (\Gamma^{(1)\mathrm{div}})^F_\Gamma &=& - g^2 \left(\zeta^{(1)}+\zeta_\lambda^{(1)}+\frac{3}{2}Z_F^{(1)}\right) _{\Gamma\Gamma'}^{FF'}B_{\Gamma'}^{F'}\,. \end{eqnarray} The $\lambda$--dependent contribution from ${\bf T}_1$ in (\ref{eq:T1}) is diagonal and given by \begin{eqnarray} g^2(\Gamma^{(1)\mathrm{div}}_\lambda)_\Gamma^F &=& 3\lambda \frac{g^2}{32 \pi^2}\frac{N+1}{N\epsilon} B_\Gamma^F\,, \end{eqnarray} so that the $\lambda$--dependent part of $\zeta$ vanishes: \begin{eqnarray} \zeta_\lambda^{(1)} &=&-\frac{3\lambda}{32 \pi^2}\frac{N+1}{N\epsilon} -\frac{3}{2} Z_F^{(1)} =\frac{\lambda}{64 N\pi^2}\frac{3(N+1)(N-3)}{\epsilon} = 0\,, \quad (N=3)\,. \end{eqnarray} Therefore $\zeta$ is $\lambda$--independent, as expected from the gauge invariance. We remark that we leave $N$ explicit in some formulae to help keep track of the origin of the various terms, but in our case we should always set $N=3$ at the end. The $\lambda$--independent part of $\Gamma^{(1)}$ from ${\bf T}_0$ in (\ref{eq:T0}) leads to $(N=3)$: \begin{eqnarray} (\Gamma^{(1)\mathrm{div}})^{fgh}_{\alpha\beta\gamma} &=&\frac{(N+1)}{2N}\frac{g^2}{16\pi^2\epsilon} \left[3 B^{fgh}_{\alpha\beta\gamma} -2B^{fgh}_{\beta\alpha\gamma} -2B^{fgh}_{\gamma\beta\alpha} -2B^{fgh}_{\alpha\gamma\beta}\right]\,, \\ (\Gamma^{(1)\mathrm{div}})^{fgh}_{\alpha\beta\hat{\gamma}} &=&\frac{(N+1)}{2N}\frac{g^2}{16\pi^2\epsilon} \left[B^{fgh}_{\alpha\beta\hat{\gamma}} -2B^{fgh}_{\beta\alpha\hat{\gamma}}\right]\,, \end{eqnarray} where $\alpha,\beta,\gamma\in\{1,2\}$ (right-handed), while $\hat\gamma\in\{\hat{1}=3,\hat{2}=4\}$ (left-handed). Note that the same results hold with hatted and unhatted indices exchanged. These relations can be easily diagonalized and the combinations which do not mix are given by \begin{eqnarray} (\zeta^{(1)} )^{fgh}_{\{\alpha\alpha\beta\}} &=& (\zeta^{(1)} )^{fgh}_{\{\hat\alpha\hat\alpha\hat\beta\}} =12\frac{d}{\epsilon}\,, \\ (\zeta^{(1)} )^{f\not=gh}_{[\alpha\beta]\alpha} &=& (\zeta^{(1)} )^{f\not=gh}_{[\hat\alpha\hat\beta]\hat\alpha} = -12\frac{d}{\epsilon}\,,\\ (\zeta^{(1)} )^{fgh}_{\{\alpha\beta\}\hat\gamma} &=& (\zeta^{(1)} )^{fgh}_{\{\hat\alpha\hat\beta\}\gamma} =4\frac{d}{\epsilon}\,, \\ (\zeta^{(1)} )^{f\not=gh}_{[\alpha\beta]\hat\gamma} &=& (\zeta^{(1)} )^{f\not=gh}_{[\hat\alpha\hat\beta]\gamma} = -12\frac{d}{\epsilon}\,, \end{eqnarray} where $d$ is given by \begin{eqnarray} d &\equiv&\frac{1}{32N\pi^2}=\frac{1}{96\pi^2}\,. \label{defd} \end{eqnarray} The square bracket denotes antisymmetrization $[\alpha\beta] =\alpha\beta - \beta\alpha$, and curly bracket means $\{\alpha\beta\} =\alpha\beta + \beta\alpha$, $\{\alpha\alpha\beta\} = \alpha\alpha\beta + \alpha\beta\alpha +\beta\alpha\alpha$. The totally symmetric case corresponds to the decuplet representation (for $N_\mathrm{f}=3$) and contains the $N_\mathrm{f}=2\,,I=3/2$ representation. The antisymmetric case corresponds to the octet representation (for $N_\mathrm{f}=3$) and contains the $N_\mathrm{f}=2\,,I=1/2$ representation. The anomalous dimension at 1-loop is easily obtained from \begin{eqnarray} \gamma &=& g^2\gamma^{(1)}+\mathrm{O}(g^4)= \beta_D(g,\epsilon)\frac{\partial\ln\zeta}{\partial g}= -2\zeta^{(1)} g^2\epsilon +\mathrm{O}(g^4)\,. \end{eqnarray} Therefore we have \begin{eqnarray} \left(\gamma^{(1)}\right)^{fgh}_{\{\alpha\alpha\beta\}} = \left(\gamma^{(1)}\right)^{fgh}_{\{\hat\alpha\hat\alpha\hat\beta\}} &=& -24d\,,\\ \left(\gamma^{(1)}\right)^{f\not=gh}_{[\alpha\beta]\alpha} =\left(\gamma^{(1)}\right)^{f\not=gh}_{[\hat\alpha\hat\beta]\hat\alpha} &=& 24d\,,\\ \left(\gamma^{(1)}\right)^{fgh}_{\{\alpha\beta\}\hat\gamma} = \left(\gamma^{(1)}\right)^{fgh}_{\{\hat\alpha\hat\beta\}\gamma} &=& -8 d\,,\\ \left(\gamma^{(1)}\right)^{f\not=gh}_{[\alpha\beta]\hat\gamma} =\left(\gamma^{(1)}\right)^{f\not=gh}_{[\hat\alpha\hat\beta]\gamma} &=& 24d\,. \end{eqnarray} \subsection{Anomalous dimensions of 6--quark operators at 1-loop} In this subsection we consider the renormalization of arbitrary local gauge invariant 6--quark operator of (lowest) dimension 9. Any such operator can be written as a linear combination of operators \begin{equation} O_C(x)=B^{F_1,F_2}_{\Gamma_1,\Gamma_2}(x)\equiv B^{F_1}_{\Gamma_1}(x) B^{F_2}_{\Gamma_2} (x)=O_A(x)O_B(x)\,, \label{sixqops} \end{equation} with $A=(\Gamma_1,F_1)$ and $B=(\Gamma_2,F_2)$. Note $O_A(x)$ and/or $O_B(x)$ may not be operators with proton or nucleon quantum numbers and separately may not be diagonally renormalizable at one loop. The reason for considering the renormalization in more generality is that in principle there may be operators in this class which occur in the OPE of two nucleon operators at higher order in PT, but are relevant in the analysis because of their potentially large anomalous dimensions. \subsubsection{Linear relations between 6--quark operators} According to the considerations in subsect.~4.2 the operators in eq.~(\ref{sixqops}) mix only with operators $O_{C'}=O_{A'}O_{B'}$ which preserve the set of flavors and Dirac indices in the chiral basis i.e. $$ F_1\cup F_2 = F'_1\cup F'_2\,,\,\,\, \Gamma_1\cup\Gamma_2 = \Gamma'_1\cup\Gamma'_2\,. $$ Note however that such operators are not all linearly independent. Relations between them follow from a general identity satisfied by the totally antisymmetric epsilon symbol which for $N$ labels reads \begin{equation} N\varepsilon^{a_1\dots a_N}\varepsilon^{b_1\dots b_N} =\sum_{j,k}\varepsilon^{a_1\dots a_{j-1}b_ka_{j+1}\dots a_N} \varepsilon^{b_1\dots b_{k-1}a_j b_{k+1}\dots b_N}\,. \end{equation} For our special case, $N=3$, this identity implies the following identities among the 6--quark operators \begin{eqnarray} 3 B^{F_1,F_2}_{\Gamma_1,\Gamma_2} + \sum_{i,j=1}^3 B^{(F_1F_2)[i,j]}_{(\Gamma_1,\Gamma_2)[i,j]} = 0\,, \label{eq:constraint} \end{eqnarray} where $i$-th index of $abc$ and $j$-th index of $def$ are interchanged in $(abc,def)[i,j]$. For example, $(\Gamma_1,\Gamma_2)[1,1]=\alpha_2\beta_1\gamma_1,\alpha_1\beta_2\gamma_2$ or $(\Gamma_1,\Gamma_2)[2,1]= \alpha_1\alpha_2\gamma_1,\beta_1\beta_2\gamma_2$. Note that the interchange of indices occurs simultaneously for both $\Gamma_1,\Gamma_2$ and $F_1,F_2$ in the above formula. The plus sign in (\ref{eq:constraint}) appears because the quark fields are Grassmann. An immediate consequence of the identity is that the divergent part of the $\lambda$--dependent contributions, calculated from ${\bf T}_1$ in (\ref{eq:1-loopT}), must vanish, after the summation over the 9 different contributions from quark pairs on the different baryonic parts $A,B$ is taken. The $\lambda$--dependent part of the contribution of quark contractions on the same baryonic parts is compensated by the quark field renormalization. Thus the renormalization of the bare 6--quark operator is $\lambda$--independent as expected from gauge invariance. As an example of identities, we consider the case that $\Gamma_1,\Gamma_2=\alpha\alpha\beta,\alpha\beta\beta$ ($\alpha\not=\beta$ and $F_1,F_2=ffg,ffg$ ($f\not= g$), the constraint gives \begin{eqnarray} && 3 B_{\alpha\alpha\beta,\alpha\beta\beta}^{ffg,ffg} + (3-2)B_{\alpha\alpha\beta,\alpha\beta\beta}^{ffg,ffg} +B_{\alpha\alpha\alpha,\beta\beta\beta}^{fff,fgg} +(2-1) B_{\alpha\beta\beta,\alpha\alpha\beta}^{fgg,fff} \nonumber \\ &=& 4 B_{\alpha\alpha\beta,\alpha\beta\beta}^{ffg,ffg}+ B_{\alpha\alpha\alpha,\beta\beta\beta}^{fff,fgg}+ B_{\alpha\beta\beta,\alpha\alpha\beta}^{fgg,fff} = 0\,, \end{eqnarray} where minus signs in the first line come from the property that $B_{\Gamma_2,\Gamma_1}^{F_2,F_1}= - B _{\Gamma_1,\Gamma_2}^{F_1,F_2}$. There are no further relations among 6--quark operators beyond (\ref{eq:constraint}). \subsubsection{Divergent parts at 1-loop} We thus need only to calculate the contributions from ${\bf T}_0$, which can be classified into the following 4 different combinations for a pair of two indices: \begin{eqnarray} \left(^f_\alpha\right) \left(^f_\alpha\right) &\Rightarrow& -(N+1) \left(^f_\alpha\right) \left(^f_\alpha\right)\,,\\ \left(^f_\alpha\right) \left(^f_\beta\right) &\Rightarrow& (1-2N) \left(^f_\alpha\right) \left(^f_\beta\right) +(N-2)\left(^f_\beta\right) \left(^f_\alpha\right)\,,\\ \left(^{f_1}_\alpha\right) \left(^{f_2}_\alpha\right) &\Rightarrow& -\left(^{f_1}_\alpha\right) \left(^{f_2}_\alpha\right) -N\left(^{f_2}_\alpha\right) \left(^{f_1}_\alpha\right)\,,\\ \left(^{f_1}_\alpha\right) \left(^{f_2}_\beta\right) &\Rightarrow& \left\{\left(^{f_1}_\alpha\right) \left(^{f_2}_\beta\right) -2\left(^{f_1}_\beta\right) \left(^{f_2}_\alpha\right) \right\} +N \left\{\left(^{f_2}_\beta\right) \left(^{f_1}_\alpha\right) -2\left(^{f_2}_\alpha\right) \left(^{f_1}_\beta\right) \right\}\,, \end{eqnarray} where $f\not= g$ and $\alpha\not= \beta \in (1,2)$ (Right) or $\in (3,4)$ (Left). The computation can be made according to the following steps: i.) Select the total flavor content e.g. $3f+3g$ or $4f+2g$ ($f\ne g$). These are the only cases we will consider since in this paper we are mainly restricting attention to baryon operators with $N_\mathrm{f}=2$, but the approach is also applicable to more general cases ($N_\mathrm{f}>2$). ii.) Given a flavor content classify all the possible sets of Dirac labels in the chiral basis e.g. $111223,112234,...$ It is obvious from the rules above that some have equivalent renormalization at 1-loop e.g. $111223$ and $112223$ with $1\leftrightarrow2$, and also those with hatted and unhatted indices exchanged e.g. $111223$ and $133344$. iii.) For given flavor and Dirac sets generate all possible operators \footnote{recall the single baryon operators are symmetric under exchange of pairs of indices}. Then generate all gauge identities between them and determine a maximally independent (basis) set $\{\mathcal{S}_i\}$. iv.) Compute the divergent parts of the members of the independent basis: \be \Gamma_i^{\mathrm{div}}=\frac{1}{2\epsilon}\gamma_{ij}\mathcal{S}_j\,. \end{equation} v.) Finally compute the eigenvalues and corresponding eigenvectors of $\gamma^T$ to determine the operators which renormalize diagonally at 1-loop. An example of the procedure is given in Appendix~A. Some of the steps are quite tedious if carried out by hand. e.g. in the case $3f+3g$ and Dirac indices $112234$ there are initially 68 operators in step iii. with 38 independent gauge identities, and hence an independent basis of 30 operators. However all the steps above can be easily implemented in an algebraic computer program using MATHEMATICA or MAPLE. If the quarks $f,g$ belong to an iso-doublet e.g. we identify $f$ with $u$ having $I_3=1/2$ and $g$ with $d$ ($I_3=-1/3$), then if an eigenvalue is non-degenerate the corresponding eigenvector belongs to a certain representation of the isospin group. If the eigenvalue is degenerate then linear combinations of them belong to definite representations. For the $3f+3g$ case they can have $I=0,1,2,3$. Eigenvectors with $I=0,2$ are odd under the interchange $f\leftrightarrow g$ and those with $I=1,3$ are even. The operators in the case $4f+2g$ have $I_3=1$ and hence have $I=1,2,3$. The eigenvectors in this case can be obtained from those of the $3f+3g$ case by applying the isospin raising operator. The complete list of eigenvalues and possible isospins are given in Tables~\ref{T3f3ga}-\ref{T3f3gc} in Appendix~A. Here we summarize the most important results. 1) For the $3f+3g$ (and $4f+2g$) cases all eigenvalues $\gamma_j\le 48d=2\gamma_N$, where $\gamma_N$ is the 1-loop anomalous dimension of the nucleon (3--quark) operator. We have not found an elegant way of proving this other than computing all cases explicitly. 2) It is easy to construct eigenvectors with eigenvalue $2\gamma_N$ e.g. operators of the form $B^{ffg}_{\alpha[\beta,\alpha]} B^{ggf}_{\hat{\alpha}[\hat{\beta},\hat{\alpha]}}$ since there is no contribution from diagrams where the gluon line joins quarks in the different baryonic parts. 3) Operators with higher isospin generally have smaller eigenvalues. \subsubsection{Decomposition of two-nucleon operators} Since \begin{eqnarray} C \gamma_5 &=& \left(\begin{array}{cccc} 0 & -1 & 0 & 0 \\ 1 & 0 & 0 & 0 \\ 0 & 0 & 0 & -1 \\ 0 & 0 & 1 & 0 \\ \end{array} \right)\,, \end{eqnarray} in the chiral representation, the nucleon operator is written as \begin{eqnarray} B_{\alpha}^f &=& B_{\alpha+\hat\alpha, [2,1]}^{ffg} +B_{\alpha+\hat\alpha, [\hat 2,\hat 1]}^{ffg} \end{eqnarray} where $\alpha = 1,2$, $\hat \alpha =\alpha+2$, and $f\not= g$. This has anomalous dimension $\gamma_N=24d$. We then consider two independent 6--quark operators occurring in the OPE at tree level which decomposed as follows. The spin-singlet ($S=0$) and isospin-triplet ($I=1$) operator is decomposed as \begin{eqnarray} B^{ffg}_{\alpha+ \hat\alpha, [\beta,\alpha]+ [\hat \beta,\hat \alpha]} B^{ffg}_{\beta+ \hat\beta, [\beta,\alpha]+ [\hat \beta,\hat \alpha]} &=& B_I^{01} + B_{II}^{01} + B_{III}^{01} + B_{IV}^{01} + B_V^{01} + B_{VI}^{01} \end{eqnarray} where \begin{eqnarray} B_I^{01} &=&B^{ffg}_{\alpha [\beta,\alpha]}B^{ffg}_{\beta [\beta,\alpha]} + B^{ffg}_{\hat\alpha [\hat\beta,\hat\alpha]}B^{ffg}_{\hat\beta[\hat\beta,\hat\alpha]}\,,\\ B_{II}^{01} &=& B^{ffg}_{\alpha [\beta,\alpha]}B^{ffg}_{\beta [\hat\beta,\hat\alpha]} +B^{ffg}_{\alpha [\hat\beta,\hat\alpha]}B^{ffg}_{\beta [\beta,\alpha]} +B^{ffg}_{\hat\alpha [\hat\beta,\hat\alpha]}B^{ffg}_{\hat\beta [\beta,\alpha]} +B^{ffg}_{\hat\alpha [\beta,\alpha]}B^{ffg}_{\hat\beta[\hat\beta,\hat\alpha]}\,, \\ B_{III}^{01} &=& B^{ffg}_{\alpha [\beta,\alpha]}B^{ffg}_{\hat\beta [\beta,\alpha]} +B^{ffg}_{\hat \alpha [\beta,\alpha]}B^{ffg}_{\beta [\beta,\alpha]} +B^{ffg}_{\hat\alpha [\hat\beta,\hat\alpha]}B^{ffg}_{\beta [\hat\beta,\hat\alpha]} + B^{ffg}_{\alpha [\hat\beta,\hat\alpha]}B^{ffg}_{\hat\beta[\hat\beta,\hat\alpha]}\,, \\ B_{IV}^{01} &=&B^{ffg}_{\alpha [\hat\beta,\hat\alpha]}B^{ffg}_{\beta [\hat\beta,\hat\alpha]} +B^{ffg}_{\hat\alpha [\beta,\alpha]}B^{ffg}_{\hat\beta[\beta,\alpha]}\,,\\ B_V^{01} &=&B^{ffg}_{\alpha [\hat\beta,\hat\alpha]}B^{ffg}_{\hat\beta [\beta,\alpha]} +B^{ffg}_{\hat\alpha [\beta,\alpha]}B^{ffg}_{\beta[\hat\beta,\hat\alpha]}\,,\\ B_{VI}^{01} &=&B^{ffg}_{\alpha [\beta,\alpha]}B^{ffg}_{\hat\beta [\hat\beta,\hat\alpha]} +B^{ffg}_{\hat\alpha [\hat\beta,\hat\alpha]}B^{ffg}_{\beta[\beta,\alpha]}\,, \end{eqnarray} where $\alpha\not= \beta$. In the above we do not have to calculate all contributions. Some of them are obtained from interchanges under $(1,2)\leftrightarrow (3,4)$ or $(1,3)\leftrightarrow (2,4)$. Similarly the spin-triplet ($S=1$) and isospin-singlet ($I=0$) operator is decomposed as \begin{eqnarray} B^{ffg}_{\alpha+ \hat\alpha, [\beta,\alpha]+ [\hat \beta,\hat \alpha]} B^{ggf}_{\alpha+ \hat\alpha, [\beta,\alpha]+ [\hat \beta,\hat \alpha]} &=& B_I^{10} + B_{II}^{10} + B_{III}^{10} + B_{IV}^{10} + B_V^{10} + B_{VI}^{10}\,, \end{eqnarray} where \begin{eqnarray} B_I^{10}&=& B^{ffg}_{\alpha [\beta,\alpha]}B^{ggf}_{\alpha [\beta,\alpha]} + B^{ffg}_{\hat\alpha [\hat\beta,\hat\alpha]}B^{ggf}_{\hat\alpha[\hat\beta,\hat\alpha]}\,, \\ B_{II}^{10}&=& B^{ffg}_{\alpha [\beta,\alpha]}B^{ggf}_{\alpha [\hat\beta,\hat\alpha]} +B^{ffg}_{\alpha [\hat\beta,\hat\alpha]}B^{ggf}_{\alpha [\beta,\alpha]} +B^{ffg}_{\hat\alpha [\hat\beta,\hat\alpha]}B^{ggf}_{\hat\alpha[\beta,\alpha]} +B^{ffg}_{\hat\alpha [\beta,\alpha]}B^{ggf}_{\hat\alpha[\hat\beta,\hat\alpha]}\,, \\ B_{III}^{10}&=& B^{ffg}_{\alpha [\beta,\alpha]}B^{ggf}_{\hat\alpha [\beta,\alpha]} +B^{ffg}_{\hat \alpha [\beta,\alpha]}B^{ggf}_{\alpha [\beta,\alpha]} +B^{ffg}_{\hat\alpha [\hat\beta,\hat\alpha]}B^{ggf}_{\alpha [\hat\beta,\hat\alpha]} +B^{ffg}_{\alpha [\hat\beta,\hat\alpha]}B^{ggf}_{\hat\alpha[\hat\beta,\hat\alpha]}\,, \\ B_{IV}^{10}&=& B^{ffg}_{\alpha [\hat\beta,\hat\alpha]}B^{ggf}_{\alpha [\hat\beta,\hat\alpha]} +B^{ffg}_{\hat\alpha [\beta,\alpha]}B^{ggf}_{\hat\alpha [\beta,\alpha]}\,, \\ B_V^{10}&=& B^{ffg}_{\alpha [\hat\beta,\hat\alpha]}B^{ggf}_{\hat\alpha [\beta,\alpha]} +B^{ffg}_{\hat\alpha [\beta,\alpha]}B^{ggf}_{\alpha[\hat\beta,\hat\alpha]}\,, \\ B_{VI}^{10}&=& B^{ffg}_{\alpha [\beta,\alpha]}B^{ggf}_{\hat\alpha [\hat\beta,\hat\alpha]} +B^{ffg}_{\hat\alpha [\hat\beta,\hat\alpha]}B^{ggf}_{\alpha[\beta,\alpha]}\,. \end{eqnarray} Again a half of the above 1-loop contributions can be obtained from others by the interchange $(1,2)\leftrightarrow (3,4)$ or $f\rightarrow g$. \subsubsection{Results for anomalous dimensions} It is very important to note here that operators $B_{VI}^{SI}$ for both cases ($SI =01$ and $10$) have the maximal anomalous dimension at 1-loop, since as noted in point 2) above, no 1-loop correction from ${\bf T}_0$ joining quarks from the two baryonic components exists for $B_{\alpha\beta\gamma,\hat\alpha'\hat\beta'\hat\gamma'}^{F_1,F_2}$ type of operators. Therefore we always have some operators with $\beta_{A} = 0$ which dominate in the OPE at short distance. The 1-loop corrections $\Gamma^{(1)}$ to 6--quark operators $B^{SI}$ are computed in appendix~\ref{appendixA} and are summarized as: \begin{eqnarray} \left(\Gamma_{I}^{01}\right)^{(1)} &=& -12\frac{d}{\epsilon}B_I^{01}\,,\quad \left(\Gamma_{II}^{01}\right)^{(1)} = 12\frac{d}{\epsilon}B_{II}^{01}\,,\quad \left(\Gamma_{III}^{01}\right)^{(1)} = 0\,,\nonumber \\ \left(\Gamma_{IV}^{01}\right)^{(1)} &=& 0\,, \ \left(\Gamma_{V}^{01}\right)^{(1)} = 6\frac{d}{\epsilon} B_{V}^{01} + 6\frac{d}{\epsilon} B_{VI}^{01}\,,\ \left(\Gamma_{VI}^{01}\right)^{(1)} = 24\frac{d}{\epsilon}B_{VI}^{01}\,, \end{eqnarray} for $SI = 01$. The last two results can be written as \begin{eqnarray} \left(\Gamma_{V'}^{01}\right)^{(1)} &=& 6\frac{d}{\epsilon} B_{V'}^{01}\,, \quad \left(\Gamma_{VI'}^{01}\right)^{(1)} = 24\frac{d}{\epsilon}B_{VI'}^{01}\,,\,, \end{eqnarray} where \begin{eqnarray} B_{V'}^{01} &=& B_V^{01} - \frac{1}{3} B_{VI}^{01}\,, \qquad B_{VI'}^{01} = B_{VI}^{01}\,. \end{eqnarray} Similarly we have for $SI=10$ \begin{eqnarray} \left(\Gamma_{I}^{10}\right)^{(1)} &=& -4\frac{d}{\epsilon}B_I^{10}\,,\quad \left(\Gamma_{II}^{10}\right)^{(1)} = 20\frac{d}{\epsilon}B_{II}^{10}\,,\quad \left(\Gamma_{III}^{10}\right)^{(1)} = 0\,,\nonumber \\ \left(\Gamma_{IV}^{10}\right)^{(1)} &=& 8\frac{d}{\epsilon} B_{IV}^{10}\,,\quad \left(\Gamma_{V'}^{10}\right)^{(1)} = 6\frac{d}{\epsilon}B_{V'}^{10}\,,\quad \left(\Gamma_{VI'}^{10}\right)^{(1)} = 24\frac{d}{\epsilon}B_{VI'}^{10}\,, \end{eqnarray} where \begin{eqnarray} B_{V'}^{10} &=& B_V^{10} - \frac{1}{3} B_{VI}^{10}\,, \qquad B_{VI'}^{10} = B_{VI}^{10}\,. \end{eqnarray} Denoting the eigenvalues of the anomalous dimension matrix by $\gamma_C$, we give the values of $\gamma^{SI}$ defined by \begin{equation} \gamma_C-2\gamma_N= 2d \gamma^{SI}\,, \label{gammaSI} \end{equation} in table~\ref{tab:gamma} ($N=3$), which shows, in both cases, that the largest value is zero while others are all negative. The case 2 in sect.~\ref{sec:3dim} is realized: $\beta_C = 0$ and \begin{eqnarray} \beta_{C^\prime} =\beta^{01}_0 &=& -\frac{6}{33-2N_\mathrm{f}} \qquad \mbox{for} \ S=0\,, I=1\,, \label{beta01}\\ \beta_{C^\prime} =\beta^{10}_0 &=& -\frac{2}{33-2N_\mathrm{f}} \qquad \mbox{for} \ S=1\,, I=0\,.\label{beta10} \end{eqnarray} \begin{table}[tb] \caption{The value of $\gamma^{SI}$ (defined in (4.72)) for each eigen operator in the $SI=01$ and $SI=10$ states. } \label{tab:gamma} \begin{center} \begin{tabular}{|c|cccccc|} \hline & $I$ & $II$ & $ III$ &$ IV$ & $V'$ & $VI'$ \\ \hline $\gamma^{01}$ & $-36$ & $-12$ & $-24$ & $-24$ & $-18$ & $0$ \\ $\gamma^{10}$ & $-28$ & $-4$ & $-24$ & $-16$ & $-18$ & $0$ \\ \hline \end{tabular} \end{center} \end{table} \section{Short distance behavior of the nucleon potential} We consider the following structure of the potential. \begin{eqnarray} V(\vec y) &=& V_0(r) + V_\sigma(r) ({\vec \sigma_1\cdot\vec \sigma_2}) + V_T(r) S_{12} +\mathrm{O}(\nabla) \end{eqnarray} where $r=\vert \vec y\vert$, and \begin{eqnarray} S_{12} &=& 3 ({\vec\sigma_1\cdot \hat{\vec y}}) ({\vec\sigma_2\cdot \hat{\vec y}}) - ({\vec \sigma_1\cdot\vec\sigma_2}), \quad \hat {\vec y} = \frac{\vec y}{\vert \vec y \vert} \label{S_12} \end{eqnarray} is the tensor operator. Here $\vec\sigma_i$ acts on the spin labels of the $i^{\rm th}$ nucleon. Since, as shown in the previous section, 6--quark operators appeared at tree level in the OPE of NN which have the largest and the second largest anomalous dimensions, we mainly consider 6--quark operators at tree level in the OPE, which is written as \begin{eqnarray} B_\alpha^f (x+y/2) B_\beta^g(x-y/2) &\simeq& c_{VI} B_{VI,\alpha\beta}^{fg}(x) + c_{II} (-\log r)^{\beta^{SI}_0} B_{II,\alpha\beta}^{fg}(x) + \cdots \label{eq:ope_full} \end{eqnarray} where $c_{VI}$ and $c_{II}$ are some constants, and $\cdots$ represents other contributions, which are less singular than the first two at short distance. (We here write the spinor and flavor indices $\alpha, \beta$ and $f,g$ explicitly for later use.) The anomalous dimensions $\beta_0^{SI}$ are given in (\ref{beta01}) and (\ref{beta10}). \subsection{Potential for $S=0$ and $I=1$ states} In the case that $S=0$ and $I=1$, we take $\alpha\not=\beta$ and $f=g$ in eq.(\ref{eq:ope_full}), whose leading contributions couple only to the $J=L=0$ state, which is given by \begin{eqnarray} \vert E \rangle = \vert L_z=0, S_z=0,I_z=1\rangle_{L=0,S=0,I=1} = \vert 0, 0,1\rangle_{0,0,1}\,. \end{eqnarray} The relevant matrix elements are given by \begin{eqnarray} c_{VI}\langle 0 \vert B_{VI,\alpha\beta}^{fg}\vert 0, 0,1\rangle_{0,0,0} &=& A_{VI}^0 Y_0^0 [\alpha\beta]\{fg\}_1,\\ c_{II}\langle 0 \vert B_{II,\alpha\beta}^{fg}\vert 0, 0,1\rangle_{0,0,0} &=& A_{II}^0 Y_0^0 [\alpha\beta]\{fg\}_1, \end{eqnarray} where $A_{II}^0$ and $A_{VI}^0$ are non-perturbative constants, $Y_L^{L_z}$ is a spherical harmonic function, $[\alpha\beta] =(\delta_{\alpha 1}\delta_{\beta 2}-\delta_{\beta 1}\delta_{\alpha 2})/\sqrt{2}$ represents the $(S,S_z)=(0,0)$ component, and $\{fg\}_1 = \delta_{f1}\delta_{g1}$ corresponds to isospin $(I,I_z) =(1,1)$. With the notation that $\phi(^1S_0,J_z=0)^{II_z=11} = Y_0^0 [\alpha\beta]\{fg\}_1$, the wave function at short distance is dominated by \begin{eqnarray} \varphi_E^{^1S_0} (y) &=&\langle 0 \vert B_\alpha^f (x+y/2) B_\beta^g(x-y/2) \vert 0,0,1\rangle_{0,0,1}\nonumber \\ &\simeq& \left( A_{VI}^0 + A_{II}^0(-\log r)^{\beta^{01}_0}\right) \phi(^1S_0, 0)^{11} +\cdots \end{eqnarray} from which we obtain \begin{eqnarray} \frac{\nabla^2}{2m} \varphi_E^{ ^1S_0} (y) &\simeq& \frac{(-\log r)^{\beta^{01}_0-1}}{r^2} \frac{-\beta^{01}_0 A_{II}^0 }{m_N} \phi(^1S_0, 0)^{11} +\cdots \end{eqnarray} where $m=m_N/2$ is the reduced mass of the two nucleon system. Since $S_{12}$ is zero on $\phi(^1S_0, 0)^{11} $, we have \begin{eqnarray} V_c^{01}(r) \varphi_E^{^1S_0} (y) &\simeq & \frac{(-\log r)^{\beta^{01}_0-1}}{r^2} \frac{-\beta^{01}_0 A_{II}^0 }{m_N} \phi(^1S_0, 0)^{11} +\cdots \end{eqnarray} where $V_c^{01}(r) = V_0(r) -3 V_\sigma (r)$. We therefore obtain \begin{eqnarray} V_c^{01}(r)&\simeq& F^{01}(r)\frac{A_{II}^0}{ A_{VI}^0}\,. \end{eqnarray} where \begin{eqnarray} F^{SI}(r) &=& \frac{-\beta^{SI}_0 (-\log r)^{\beta^{SI}_0-1}}{m_N r^2}\,. \end{eqnarray} The potential diverges as $F^{01}(r)$ in the $r\rightarrow 0$ limit, which is a little weaker than $r^{-2}$. \subsection{Potential for $S=1$ and $I=0$} We here consider the spin-triplet and isospin-singlet state($S=1$ and $I=0$). Since $I=0$ and $I_z=0$ in this case, we drop indices $I$ and $I_z$ unless necessary. In this case, the leading contributions in eq.(\ref{eq:ope_full}) couple only to the $J=1$ state, which is given by \begin{eqnarray} \vert E \rangle &=& \vert ^3S_1,J_z=1\rangle + x \vert^3D_1,J_z=1\rangle\,, \end{eqnarray} where \begin{eqnarray} \vert ^3 S_1, J_z=1\rangle &=& \vert L_z=0, S_z=1\rangle_{L=0,S=1}\,,\\ \vert ^3 D_1, J_z=1\rangle &=& \frac{1}{\sqrt{10}}\left[ \vert 0, 1\rangle -\sqrt{3} \vert 1,0\rangle + \sqrt{6}\vert 2, -1\rangle\right]_{L=2,S=1}\,, \end{eqnarray} $x$ is the mixing coefficient, which is determined by QCD dynamics, and $^{2S+1} L_J$ specifies quantum numbers of the state. Relevant matrix elements are given by \begin{eqnarray} c_i \left\langle 0 \left\vert B_{i,\alpha\beta}^{fg} \right\vert {}^3S_1, 1\right\rangle &=& B_i^0 \phi\left({}^3S_1\right)\,, \qquad c_i \left\langle 0\left\vert B_{i,\alpha\beta}^{fg} \right\vert {}^3D_1, 1\right\rangle = 0\,, \end{eqnarray} for $i=II$ and $VI$, where $B_i^0$ are non-perturbative constants, and \begin{eqnarray} \phi\left({}^3S_{1}\right) &=& Y_0^0(\theta,\phi)\{\alpha\beta\}_1[fg]\,, \end{eqnarray} with $\{\alpha\beta\}_1=\delta_{\alpha 1}\delta_{\beta 1}$. Using the above results, we have \begin{eqnarray} \varphi_E^{J=1}(y) &\simeq& \left\{ B_{VI}^0 + (-\log r)^{\beta^{10}_{0}} B_{II}^0\right\} \phi\left({}^3S_{1}\right) , \end{eqnarray} By applying $\nabla^2$ we obtain \begin{eqnarray} \frac{\nabla^2}{2m}\varphi_E^{J=1} (y) &\simeq& \frac{-\beta^{10}_{}}{m_N} \frac{(-\log r)^{\beta^{10}_{0}-1}}{ r^2} B_{II}^0\phi\left({}^3S_{1}\right) \,, \label{eq:derV} \end{eqnarray} while \begin{eqnarray} V(y) \varphi_E^{J=1}(y) &\simeq& B_{VI}^0 V_c^{10}(r) \phi\left({}^3S_{1}\right) + 2\sqrt{2} V_T(r) B_{VI}^0 \phi\left({}^3D_{1}\right)\,, \label{eq:genV} \end{eqnarray} where we use the formula in the appendix~\ref{appendixB}, $V_c^{10}(r) = V_0^0(r) + V_\sigma^0 (r)$, and \begin{eqnarray} \phi\left({}^3D_{1}\right) &=& \frac{1}{\sqrt{10}}\left[ Y_{2}^0\{\alpha\beta\}_1-\sqrt{3} Y_{2}^{1}\{\alpha\beta\}_0 +\sqrt{6}Y_{2}^{2}\{\alpha\beta\}_{-1}\right] [fg]\,, \end{eqnarray} with $\{\alpha\beta\}_{-1}=\delta_{\alpha 2}\delta_{\beta 2}$ and $\{\alpha\beta\}_0=(\delta_{\alpha 1}\delta_{\beta 2}+\delta_{\beta1}\delta_{\alpha 2})/\sqrt{2}$. By comparing eq.(\ref{eq:derV}) with eq.(\ref{eq:genV}), we have \begin{eqnarray} V_c^{10}(r) &\simeq& F^{10}(r) \frac{B_{II}^0}{B_{VI}^0}\,, \qquad V_T(r) \simeq 0\,. \end{eqnarray} This shows that the central potential $V_c^{10}(r)$ diverges as $F^{10}(r)$ in the $r\rightarrow 0$ limit, which is a little weaker than $1/r^2$, while the tensor potential $V_T(r)$ becomes zero in this limit at the tree level in the OPE. \subsection{Higher order in the OPE and the tensor potential} While the short distance behavior of the central potential is determined by the OPE at the tree level, the determination of the tensor potential at short distance requires the OPE at higher order, whose relevant contribution is given by \begin{eqnarray} B_\alpha^f (x+y/2) B_\beta^g(x-y/2) &\simeq& c_{VI} B_{VI,\alpha\beta}^{fg}(x) + c_{II} (-\log r)^{\beta^{SI}_0} B_{II,\alpha\beta}^{fg}(x) \nonumber \\ &+& c_T (-\log r)^{\beta^{SI}_T} \frac{[ y^k y^l]}{r^2}B^{fg,kl}_{T,\alpha\beta}(x) +\cdots \cdots \label{eq:ope_high} \end{eqnarray} where $[y^ky^l] = y^k y^l - r^2\delta^{kl}/3$, and we assume that the third term with the tensor-type operator $B_{T,\alpha\beta}^{fg,kl}$ appears first at $\ell_T^{SI} ( > 0) $ loop of the perturbative expansion in the OPE. Therefore, with this assumption, we have \begin{eqnarray} \beta_T^{SI} &=& - \ell_T^{SI} + \frac{\Delta_T^{SI} - 24}{2(33-2N_f)} \end{eqnarray} where $\Delta^{SI}_T =\gamma_T/(2d)$ with $\gamma_T$ being the anomalous dimension of the operator $B_{T,\alpha\beta}^{fg,kl}$. The calculation of anomalous dimensions for all 6--quark operators in the previous section shows that $\Delta_T \le 24$, so that $ \beta_T^{SI} < \beta^{SI}_0 < 0$. An extra matrix element we need is given as \begin{eqnarray} x\, c_T \frac{[ y^ky^l]}{r^2} \left\langle 0\left\vert B_{T,\alpha\beta}^{fg,kl} \right\vert {}^3D_{1}, 1\right\rangle &=& B_T \phi\left({}^3D_{1}\right)\,, \end{eqnarray} where $B_T$ is a further non-perturbative constant. Using the above results, we have for $(S,I) = (1,0)$ \begin{eqnarray} \varphi_E^{J=1}(y) &\simeq& \left\{ B_{VI}^0 + (-\log r)^{\beta^{10}_{0}} B_{II}^0\right\} \phi\left({}^3S_{1}\right) + (-\log r)^{\beta^{10}_T} B_T \phi\left({}^3D_{1}\right)\,. \end{eqnarray} By applying $\nabla^2$ we obtain \begin{eqnarray} \frac{\nabla^2}{2m}\varphi_E^{J=1} (y) &\simeq& \frac{-\beta^{10}_{0}}{m_N} \frac{(-\log r)^{\beta^{10}_{0}-1}}{ r^2} B_{II}^0\phi\left({}^3S_{1}\right) \nonumber \\ &+& \frac{-6}{m_N} \frac{(-\log r)^{\beta^{10}_{T}}}{ r^2} B_{T} \phi\left({}^3D_{1}\right)\,. \end{eqnarray} From $V(y)\varphi_E^{J=1}(y) = (E + \nabla^2/(2m) )\varphi_E^{J=1}(y) $, we obtain \begin{eqnarray} V_c^{10}(r) &\simeq& F^{10}(r) \frac{B_{II}^0}{B_{VI}^0}\,, \\ V_T(r) &\simeq& F_{T}^{10}(r) \frac{-3 B_{T}}{\sqrt{2} B_{VI}^0}\,, \label{eq:eq2a} \end{eqnarray} where \begin{eqnarray} F_T^{10}(r) &=& \frac{(-\log r)^{\beta^{10}_T}}{m_N r^2}\,. \end{eqnarray} This shows that the central potential $V_c(r)$ diverges as $F^{10}(r)$ in the $r\rightarrow 0$ limit, which is a little weaker than $1/r^2$, while the tensor potential $V_T(r)$ diverges as $F^{10}_T(r)$ in this limit, which is not stronger than $F^{10}(r)$ since $\beta^{10}_0 -1 \ge \beta^{10}_T$. \section{Evaluation of matrix elements} We rewrite 3--quark operators in terms of left- and right- handed component: \begin{eqnarray} B_{X,\alpha}^f &=& B_{\alpha\beta\gamma}^{fgh} (C\gamma_5 P_X)_{\beta\gamma}(i\tau_2)_{gh}\,, \\ B_{XY,\alpha}^f &=& (P_X)_{\alpha\beta} B_{Y,\beta}^f\,, \end{eqnarray} for $X,Y = R$ or $L$. In terms of these we have \begin{eqnarray} B_{\alpha}^f B_{\beta}^g &=& \left[B_I + B_{II} + B_{III} + B_{IV} + B_{V} + B_{VI}\right]^{fg}_{\alpha\beta}\,, \end{eqnarray} where \begin{eqnarray} (B_I)^{fg}_{\alpha\beta} &=& \left[ B_{RR}B_{RR}+B_{LL}B_{LL}\right]^{fg}_{\alpha\beta}\,,\\ (B_{II})^{fg}_{\alpha\beta} &=& \left[ B_{RR}B_{RL}+B_{RL}B_{RR} +B_{LL}B_{LR}+B_{LR}B_{LL}\right]^{fg}_{\alpha\beta}\,,\\ (B_{III})^{fg}_{\alpha\beta} &=& \left[ B_{RR}B_{LR} +B_{LR}B_{RR}+B_{LL}B_{RL}+B_{RL}B_{LL}\right]^{fg}_{\alpha\beta}\,,\\ (B_{IV})^{fg}_{\alpha\beta} &=& \left[ B_{RL}B_{RL}+B_{LR}B_{LR}\right]^{fg}_{\alpha\beta}\,,\\ (B_{V})^{fg}_{\alpha\beta} &=& \left[ B_{RL}B_{LR} +B_{LR}B_{RL}\right]^{fg}_{\alpha\beta}\,,\\ (B_{VI})^{fg}_{\alpha\beta} &=& \left[ B_{RR}B_{LL} +B_{LL}B_{RR}\right]^{fg}_{\alpha\beta}\,. \end{eqnarray} Note that we take $(\vec x , t) =(\vec 0,0)$ in the above operators. We need to know \begin{eqnarray} \langle 0 \vert (B_i)^{fg}_{\alpha\beta} \vert 2N, E\rangle \end{eqnarray} for $i=II,VI$. For $f\not= g$, Lorentz covariance leads to \begin{eqnarray} \langle 0 \vert B_X^f B_Y^g \vert 2{\rm N}, E\rangle &=& \sum_{A,B=R,L} C_{XY}^{AB}(s) P_A u (\vec p, \sigma_1) P_B u(-\vec p,\sigma_2), \label{eq:FF} \end{eqnarray} where $s= E^2 = 4 \sqrt{\vec p^2 +m_N^2}$ with the total energy $E$ in the center of mass frame, $\sigma_i$ ($i=1,2$) is the spin of the $i$-th nucleon, and $C_{XY}^{AB}$ is an unknown function of $s$. Using invariance of QCD under the parity transformation $ P B_X P^{-1} = \gamma_4 B_{\bar X} $ where $\bar R = L$ and $\bar L = R$, we rewrite eq.(\ref{eq:FF}) as \begin{eqnarray} (\ref{eq:FF}) &=& \langle 0 \vert P B_X^f B_Y^g P^{-1} P \vert 2{\rm N}, E\rangle = \sum_{A.B} C_{\bar X \bar Y}^{AB} P_{\bar A} \gamma_4 u(-\vec p,\sigma_1) P_{\bar B} \gamma_4 u (\vec p, \sigma_2) \nonumber \\ &=& \sum_{A,B} C_{\bar X \bar Y}^{\bar A\bar B} P_{A} u(\vec p,\sigma_1) P_{B} u (-\vec p, \sigma_2), \end{eqnarray} where $\gamma_4 u(-\vec p,\sigma_1) = u(\vec p,\sigma_1)$ is used. The above relation implies $C_{\bar X \bar Y}^{\bar A\bar B} = C_{XY}^{AB}$. Using this property for the unknown functions $C_{XY}^{AB}$, we have \begin{eqnarray} \langle 0 \vert (B_{II})^{fg\pm gf}_{\alpha\beta} \vert 2{\rm N}, E\rangle &=& C_{RL+LR}^{RR,\pm}\{ (P_R\otimes P_R + P_L\otimes P_L) u(\vec p,\sigma_1) u(-\vec p,\sigma_2) \}_{\alpha\beta\mp\beta\alpha} \end{eqnarray} and \begin{eqnarray} \langle 0 \vert (B_{VI})^{fg\pm gf}_{\alpha\beta} \vert 2{\rm N}, E\rangle &=& C_{RL}^{RL,\pm}\{ (P_R\otimes P_L + P_L\otimes P_R) u(\vec p,\sigma_1) u(-\vec p,\sigma_2) \}_{\alpha\beta\mp\beta\alpha} \end{eqnarray} Taking $\vec p =(0,0, p_z> 0)$ and Dirac representation for $\gamma$ matrices \cite{IZ}, we have \begin{eqnarray} u(\pm\vec p, +) &= \dfrac{1}{\sqrt{E_N+ m_N}}\left( \begin{array}{c} E_N + m_N \\ 0 \\ \mp p_z \\ 0 \\ \end{array} \right)\, u(\pm \vec p, -) &= \frac{1}{\sqrt{E_N+ m_N}}\left( \begin{array}{c} 0 \\ E_N + m_N \\ 0 \\ \pm p_z \\ \end{array} \right) \end{eqnarray} where $E_N=\sqrt{\vec p^2+m_N^2}$. For $I=1$ ( $fg+gf$) and $ S=0$ ($ \sigma_1=+$ and $\sigma_2=-$ ) the above explicit form for the spinors gives \begin{eqnarray} \{ (P_R\otimes P_R + P_L\otimes P_L) u(\vec p,+) u(-\vec p,-) \}_{12-21} &=& E_N\,, \\ \{ (P_R\otimes P_L + P_L\otimes P_R) u(\vec p,+) u(-\vec p,-) \}_{12-21} &=& m_N\, , \end{eqnarray} while, for $I=0$ ( $fg-gf$) and $ S=1$ ($ \sigma_1=+$ and $\sigma_2=+$ ), we have \begin{eqnarray} \{ (P_R\otimes P_R + P_L\otimes P_L) u(\vec p,+) u(-\vec p,+) \}_{11} &=& m_N\,, \\ \{ (P_R\otimes P_L + P_L\otimes P_R) u(\vec p,+) u(-\vec p,+) \}_{11} &=& E_N\, . \end{eqnarray} We finally obtain \begin{eqnarray} \frac{\langle 0 \vert (B_{II})^{fg+gf}_{12} \vert 2{\rm N}, E\rangle } {\langle 0 \vert (B_{VI})^{fg+gf}_{12} \vert 2{\rm N}, E\rangle } &=& \frac{ E_N}{m_N}\frac{C_{RL+LR}^{RR,+}(s)}{C_{RL}^{RL,+}(s)} \end{eqnarray} for $fg+gf$ and $(\sigma_1,\sigma_2)=(+,-)$ ( $^1S_0$ ), and \begin{eqnarray} \frac{\langle 0 \vert (B_{II})^{fg-fg}_{11} \vert 2{\rm N}, E\rangle } {\langle 0 \vert (B_{VI})^{fg-fg}_{11} \vert 2{\rm N}, E\rangle } &=& \frac{m_N}{E_N} \frac{C_{RL+LR}^{RR,-}(s)}{C_{RL}^{RL,-}(s)} \end{eqnarray} for $fg-gf$ and $(\sigma_1,\sigma_2)=(+,+)$ ( $^3S_1$ ), where $s=4E_N^2$. Unfortunately, we can not determine the sign of the ratio for these matrix elements. As a very crude estimation, we consider the non-relativistic expansion for constituent quarks whose mass $m_Q$ is given by $m_Q=m_N/3$. In the large $m_Q$ limit, $\gamma_4 q_0 = q_0$ and $\gamma_4 u_0 = u_0$, where a subscript $0$ for $q$ and $u$ means the 0-th order in the non-relativistic expansion. In this limit, it is easy to show $C_{XY}^{AB} = C$ for all $X,Y,A,B$, so that $C^{RR}_{RL+LR}=2C$ and $C_{RL}^{RL}=C$. Furthermore the first order correction to $C_{XY}^{AB} = C$ vanishes in the expansion. Therefore in the leading order of the non-relativistic expansion, we have \begin{eqnarray} \frac{\langle 0 \vert (B_{II})^{fg+gf}_{12} \vert 2{\rm N}, E\rangle } {\langle 0 \vert (B_{VI})^{fg+gf}_{12} \vert 2{\rm N}, E\rangle } &\simeq& 2 + \mathrm{O}\left(\frac{{\vec p}^2}{m_Q^2}\right) \end{eqnarray} for $(\sigma_1,\sigma_2)=(+,-)$ ( $^1S_0$ ), and \begin{eqnarray} \frac{\langle 0 \vert (B_{II})^{fg-fg}_{11} \vert 2{\rm N}, E\rangle } {\langle 0 \vert (B_{VI})^{fg-fg}_{11} \vert 2{\rm N}, E\rangle } &\simeq& 2 + \mathrm{O}\left(\frac{{\vec p}^2}{m_Q^2}\right) \end{eqnarray} for $(\sigma_1,\sigma_2)=(+,+)$ ( $^3S_1$ ). For both cases, we have positive sign for the ratio, which gives repulsion at short distance, the repulsive core. \section{Conclusions and discussion} The OPE analysis leads to conclusion that the S--state potential at short distance behaves as in (\ref{summary1}) with (\ref{summary2}). However perturbative considerations alone can not tell the crucial sign of the overall coefficient $C_E$. Moreover we found that the latter was also not directly predicted by chiral PT . A crude estimation using non-relativistic quarks suggests that $C_E$ is positive, hence predicting a repulsive core, which diverges a little weaker than $r^{-2}$ at small $r$. The leading corrections involve small powers of logs and hence it could happen that the dominant asymptotic behavior appears only at extremely short distances. Our analysis suggests that the repulsion of the NN potential at short distance is related to the difference of anomalous dimensions between a 6--quark operator and two 3--quark operators at 1-loop, and to the structure of the composite operators which probe the NN states. The explicit 1--loop calculation indicates that a combination of fermi statistics for quarks and the particular structure of the one gluon exchange interaction determines the sign and size of the $\beta$'s. The appearance of zero effective gamma eigenvalues is simply explained by chiral symmetry, however we were unable to find a simple proof of the absence of positive eigenvalues established by explicit calculation. One speculation is that the latter intriguing pattern can be explained by the relation between the 1-loop QCD anomalous dimensions and those of super YM theory \cite{Beisert}. At higher order in the perturbative expansion, tensor operators appear in the OPE. Using this fact, we also found that the tensor potential also diverges a little weaker than $r^{-2}$ as $r\rightarrow 0$. There are several interesting extensions of the analysis using the OPE. An application to the 3--flavor case may reveal the nature of the repulsive core in the baryon-baryon potentials. Since quark masses can be neglected in our OPE analysis, the calculation can be done in the exact SU(3) symmetric limit. It is also interesting to investigate the existence or the absence of the repulsive core in the 3--body nucleon potential. Such an investigation would require the calculation of anomalous dimensions of 9--quark operators at 2--loop level. Certainly more precise evaluations (also involving numerical simulations) of matrix elements $\langle 0 \vert O_X\vert E\rangle$ will also be needed to theoretically predict the nature of the core of the NN potential. \section*{Acknowledgments} S.~A would like to thank Dr. T.~Doi, Prof. T.~Hatsuda, Dr. N.~Ishii and Prof. W.~Wise for useful discussions. S.~A. is supported in part by Grant-in-Aid of the Ministry of Education, Sciences and Technology, Sports and Culture (Nos. 20340047, 20105001, 20105003). S.~A. and J.~B. would like to thank the Max-Planck-Institut f\"ur Physik for its kind hospitality during our stay for this research project.
\section{Introduction} \label{s:Introduction} Vector auto--regressive (VAR) models play an important role in contemporary macro--economics, being an example of an approach called the ``dynamic stochastic general equilibrium'' (DSGE), which is superseding traditional large--scale macro--econometric forecasting methodologies~\cite{Sims1980}. The motivation behind them is based on the assertion that more recent values of a variable are more likely to contain useful information about its future movements than the older ones. On the other hand, a standard tool in multivariate time series analysis is vector moving average (VMA) models, which is really a linear regression of the present value of the time series w.r.t. the past values of a white noise. A broader class of stochastic processes used in macro--economy comprises both these kinds together in the form of vector auto--regressive moving average (VARMA) models. These methodologies can capture certain spatial and temporal structures of multidimensional variables which are often neglected in practice; including them not only results in more accurate estimation, but also leads to models which are more interpretable. They are widely used by academia and central banks (\emph{cf.} the European Central Bank's Smets--Wouters model for the euro zone~\cite{SmetsWouters2002}), as they constitute quite a simple version of the DSGE equations. VARMA models are constructed from a number of univariate ARMA (Box--Jenkins; see for example~\cite{BoxJenkinsReinsel1994}) processes, typically coupled with each other. In this paper, we investigate only a significantly simplified circumstance when there is no coupling between the many ARMA components. One may argue that this is too far fetched and will be of no use in describing an economic reality. However, one may also treat it as a ``zeroth--order hypothesis,'' analogously to the idea of~\cite{LalouxCizeauBouchaudPotters1999,PlerouGopikrishnanRosenowAmaralStanley1999} in finance, namely that the case with no cross--covariances is considered theoretically, and subsequently compared to some real--world data modeled by a VARMA process; any discrepancy between the two will reflect nontrivial cross--covariances present in the system, thus permitting their investigation. This latter route is taken in this communication. A challenging and yet increasingly important problem is the estimation of large covariance matrices generated by these stationary \smash{$\mathrm{VARMA} ( q_{1} , q_{2} )$} processes, since high dimensionality of the data as compared to the sample size is quite common in many statistical problems (the ``dimensionality curse''). Therefore, an appropriate ``noise cleaning'' procedure has to be implemented, and random matrix theory (RMT) provides a natural and efficient outfit for doing that. In particular, the mean spectral densities (a.k.a. ``limiting spectral distributions,'' LSD) of the Pearson estimators of the cross--covariances for the $\mathrm{VMA} ( 1 )$ and $\mathrm{VAR} ( 1 )$ models, in the relevant high--dimensionality sector and under the full decoupling, have been derived in~\cite{JinWangMiaoHuang2009} by applying the framework proposed by~\cite{BaiSilverstein2006}. In this paper, we suggest that such calculations can be considerably simplified by resorting to a mathematical concept of the free random variables (FRV) calculus~\cite{VoiculescuDykemaNica1992,Speicher1994}, succinctly introduced in sec.~\ref{s:DoublyCorrelatedWishartEnsemblesAndFreeRandomVariables}. Our general FRV formula~\cite{BurdaJaroszJurkiewiczNowakPappZahed2009} allows not only to rediscover, which much less strain, the two fourth--order polynomial equations obtained in~\cite{JinWangMiaoHuang2009} in the $\mathrm{VMA} ( 1 )$ and $\mathrm{VAR} ( 1 )$ cases, but also to derive a sixth--order equation (\ref{eq:VARMAOneOneMainEquation}) which produces the mean spectral density for a more involved $\mathrm{VARMA} ( 1 , 1 )$ model. The results are verified by numerical simulations, which show a perfect agreement. This is all done in sec.~\ref{s:VARMAFromFreeRandomVariables}. \section{Doubly Correlated Wishart Ensembles and Free Random Variables} \label{s:DoublyCorrelatedWishartEnsemblesAndFreeRandomVariables} \subsection{Doubly Correlated Wishart Ensembles} \label{ss:DoublyCorrelatedWishartEnsembles} \subsubsection{Correlated Gaussian Random Variables} \label{sss:CorrelatedGaussianRandomVariables} $\mathrm{VARMA} ( q_{1} , q_{2} )$ stochastic processes, as we will see below, fall within quite a general set--up encountered in many areas of science where a probabilistic nature of multiple degrees of freedom evolving in time is relevant, for example, multivariate time series analysis in finance, applied macro--econometrics and engineering. To describe this framework, consider a situation of $N$ time--dependent random variables which are measured at $T$ consecutive time moments (separated by some time interval $\delta t$); let \smash{$Y_{i a}$} be the value od the $i$--th ($i = 1 , \ldots , N$) random number at the $a$--th time moment ($a = 1 , \ldots , T$); together, they make up a rectangular $N \times T$ matrix $\mathbf{Y}$. In what usually would be the first approximation, each \smash{$Y_{i a}$} is supposed to be drawn from a Gaussian probability distribution. We will also assume that they have mean values zero, \smash{$\langle Y_{i a} \rangle = 0$}. These degrees of freedom may in principle display mutual correlations. A set of correlated zero--mean Gaussian numbers is fully characterized by the two--point covariance function, \smash{$\mathcal{C}_{i a , j b} \equiv \langle Y_{i a} Y_{j b} \rangle$} if the underlying stochastic process generating these numbers is stationary. Linear stochastic processes, including $\mathrm{VARMA} ( q_{1} , q_{2} )$, belong to this category. We will restrict our attention to an even narrower class where the cross--correlations between different variables and the auto--correlations between different time moments are factorized, \emph{i.e.}, \begin{equation}\label{eq:CovarianceFunctionFactorization} \left\langle Y_{i a} Y_{j b} \right\rangle = C_{i j} A_{a b} . \end{equation} In this setting, the inter--variable covariances do not change in time (and are described by an $N \times N$ cross--covariance matrix $\mathbf{C}$), and also the temporal covariances are identical for all the numbers (and are included in a $T \times T$ auto--covariance matrix $\mathbf{A}$; both these matrices are symmetric and positive--definite). The Gaussian probability measure with this structure of covariances is known from textbooks, $$ P_{\mathrm{c.G.}} ( \mathbf{Y} ) \mathrm{D} \mathbf{Y} = \frac{1}{\mathcal{N}_{\mathrm{c.G.}}} \exp \left( - \frac{1}{2} \sum_{i , j = 1}^{N} \sum_{a , b = 1}^{T} Y_{i a} \left[ \mathbf{C}^{- 1} \right]_{i j} Y_{j b} \left[ \mathbf{A}^{- 1} \right]_{b a} \right) \mathrm{D} \mathbf{Y} = $$ \begin{equation}\label{eq:CorrelatedGaussian} = \frac{1}{\mathcal{N}_{\mathrm{c.G.}}} \exp \left( - \frac{1}{2} \mathrm{Tr} \mathbf{Y}^{\mathrm{T}} \mathbf{C}^{- 1} \mathbf{Y} \mathbf{A}^{- 1} \right) \mathrm{D} \mathbf{Y} , \end{equation} where the normalization constant \smash{$\mathcal{N}_{\mathrm{c.G.}} = ( 2 \pi )^{N T / 2} ( \mathrm{Det} \mathbf{C} )^{T / 2} ( \mathrm{Det} \mathbf{A} )^{N / 2}$}, and the integration measure \smash{$\mathrm{D} \mathbf{Y} \equiv \prod_{i = 1}^{N} \prod_{a = 1}^{T} \mathrm{d} Y_{i a}$}, while the letters ``c.G.'' stand for ``correlated Gaussian.'' Now, a standard way to approach correlated Gaussian random numbers is to recall that they can always be decomposed as linear combinations of uncorrelated Gaussian degrees of freedom; indeed, this is achieved through the transformation \begin{equation}\label{eq:UncorrelatedGaussian} \mathbf{Y} = \sqrt{\mathbf{C}} \widetilde{\mathbf{Y}} \sqrt{\mathbf{A}} , \qquad \textrm{which yields} \qquad P_{\mathrm{G.}} ( \widetilde{\mathbf{Y}} ) \mathrm{D} \widetilde{\mathbf{Y}} = \frac{1}{\mathcal{N}_{\mathrm{G.}}} \exp \left( - \frac{1}{2} \mathrm{Tr} \widetilde{\mathbf{Y}}^{\mathrm{T}} \widetilde{\mathbf{Y}} \right) \mathrm{D} \widetilde{\mathbf{Y}} , \end{equation} where the square roots of the covariance matrices, necessary to facilitate the transition, exist due to the positive--definiteness of $\mathbf{C}$ and $\mathbf{A}$; the new normalization reads \smash{$\mathcal{N}_{\mathrm{G.}} = ( 2 \pi )^{N T / 2}$}. \subsubsection{Estimating Equal--Time Cross--Covariances} \label{sss:EstimatingEqualTimeCrossCovariances} An essential problem in multivariate analysis is to determine (estimate) the covariance matrices $\mathbf{C}$ and $\mathbf{A}$ from given $N$ time series of length $T$ of the realizations of our random variables \smash{$Y_{i a}$}. For simplicity, we do not distinguish in notation between random numbers, \emph{i.e.}, the population, and their realizations in actual experiments, \emph{i.e.}, the sample. Since the realized cross--covariance between degrees $i$ and $j$ at the same time $a$ is \smash{$Y_{i a} Y_{j a}$}, the simplest method to estimate the today's cross--covariance \smash{$c_{i j}$} is to compute the time average, \begin{equation}\label{eq:EstimatorcDefinition} c_{i j} \equiv \frac{1}{T} \sum_{a = 1}^{T} Y_{i a} Y_{j a} , \qquad \textrm{\emph{i.e.}, } \qquad \mathbf{c} = \frac{1}{T} \mathbf{Y} \mathbf{Y}^{\mathrm{T}} = \frac{1}{T} \sqrt{\mathbf{C}} \widetilde{\mathbf{Y}} \mathbf{A} \widetilde{\mathbf{Y}}^{\mathrm{T}} \sqrt{\mathbf{C}} . \end{equation} This is usually named the ``Pearson estimator'', up to the prefactor which depending on the context is $1 / ( T - 1 )$ or $1 / T$. Other estimators might be introduced, such as between distinct degrees of freedom at separate time moments (``time--delayed estimators''), or with certain decreasing weights given to older measurements to reflect their growing obsolescence (``weighted estimators''), but we will not investigate them in this article. Furthermore, in the last equality in (\ref{eq:EstimatorcDefinition}), we cast $\mathbf{c}$ through the uncorrelated Gaussian numbers contained in \smash{$\widetilde{\mathbf{Y}}$}, the price to pay for this being that the covariance matrices now enter into the expression for $\mathbf{c}$, making it more complicated; this will be the form used hereafter. The random matrix $\mathbf{c}$ is called a ``doubly correlated Wishart ensemble''~\cite{Wishart1928}. Let us also mention that the auto--covariance matrix $\mathbf{A}$ can be estimated through \smash{$\mathbf{a} \equiv ( 1 / N ) \mathbf{Y}^{\mathrm{T}} \mathbf{Y}$}. However, it is verified that this object carries identical information to the one contained in $\mathbf{c}$ (it is ``dual'' to $\mathbf{c}$), and therefore may safely be discarded. Indeed, these two estimators have same non--zero eigenvalues (modulo an overall rescaling by $r$), and the larger one has $| T - N |$ additional zero modes. Any historical estimator is inevitably marred by the measurement noise; it will reflect the true covariances only to a certain degree, with a superimposed broadening due to the finiteness of the time series. More precisely, there are $N ( N + 1 ) / 2$ independent elements in $\mathbf{C}$, to be estimated from $N T$ measured quantities $\mathbf{Y}$, hence the estimation accuracy will depend on the ``rectangularity ratio,'' \begin{equation}\label{eq:NoiseToSignalRatio} r \equiv \frac{N}{T} ; \end{equation} the closer $r$ to zero, the more truthful the estimate. This is a cornerstone of classical multivariate analysis. Unfortunately, a practical situation will typically feature a large number of variables sampled over a comparably big number of time snapshots, so that we may approximately talk about the ``thermodynamical limit,'' \begin{equation}\label{eq:ThermodynamicalLimit} N \to \infty , \qquad T \to \infty , \qquad \textrm{such that} \qquad r = \textrm{fixed} . \end{equation} On the other hand, it is exactly this limit in which the FRV calculus (see the subsection below for its brief elucidation) can be applied; hence, the challenge of de--noising is somewhat counterbalanced by the computationally powerful FRV techniques. \subsection{A Short Introduction to the Free Random Variables Calculus: The Multiplication Algorithm} \label{ss:AShortIntroductionToTheFreeRandomVariablesCalculusTheMultiplicationAlgorithm} \subsubsection{The $M$--Transform and the Spectral Density} \label{sss:TheMTransformAndTheSpectralDensity} Any study of a (real symmetric $K \times K$) random matrix $\mathbf{H}$ will most surely include a fundamental question about the average values of its (real) eigenvalues \smash{$\lambda_{1} , \ldots , \lambda_{K}$}. They are concisely encoded in the ``mean spectral density,'' \begin{equation}\label{eq:MeanSpectralDensityDefinition} \rho_{\mathbf{H}} ( \lambda ) \equiv \frac{1}{K} \sum_{i = 1}^{K} \left\langle \delta \left( \lambda - \lambda_{i} \right) \right\rangle = \frac{1}{K} \left\langle \mathrm{Tr} \left( \lambda \mathbf{1}_{K} - \mathbf{H} \right) \right\rangle . \end{equation} Here the expectation map $\langle \ldots \rangle$ is understood to be taken w.r.t. the probability measure $P ( \mathbf{H} ) \mathrm{D} \mathbf{H}$ of the random matrix. We will always have this distribution rotationally (\emph{i.e.}, \smash{$\mathbf{H} \to \mathbf{O}^{\mathrm{T}} \mathbf{H} \mathbf{O}$}, with $\mathbf{O}$ orthogonal) invariant, and hence the full information about $\mathbf{H}$ resides in its eigenvalues, distributed on average according to (\ref{eq:MeanSpectralDensityDefinition}). On the practical side, it is more convenient to work with either of the two equivalent objects, \begin{equation}\label{eq:GreenFunctionAndMTransformDefinition} G_{\mathbf{H}} ( z ) \equiv \frac{1}{K} \left\langle \mathrm{Tr} \frac{1}{z \mathbf{1}_{K} - \mathbf{H}} \right\rangle , \qquad \textrm{or} \qquad M_{\mathbf{H}} ( z ) \equiv z G_{\mathbf{H}} ( z ) - 1 , \end{equation} referred to as the ``Green's function'' (or the ``resolvent'') and the ``$M$--transform'' of $\mathbf{H}$. The latter is also called the ``moments' generating function,'' since if the ``moments'' \smash{$M_{\mathbf{H} , n} \equiv ( 1 / K ) \langle \mathrm{Tr} \mathbf{H}^{n} \rangle$} of $\mathbf{H}$ exist, it can be expanded into a power series around $z \to \infty$ as \smash{$M_{\mathbf{H}} ( z ) = \sum_{n \geq 1} M_{\mathbf{H} , n} / z^{n}$}. It should however be underlined that even for probability measures disallowing such an expansion (heavy--tailed distributions, preeminent in finance, being an example), the quantities (\ref{eq:GreenFunctionAndMTransformDefinition}) still manage to entirely capture the spectral properties of $\mathbf{H}$; hence the name ``$M$--transform'' more appropriate, in addition to being more compact. We will show that for our purposes (multiplication of random matrices; see par.~\ref{sss:TheNTransformAndFreeRandomVariables}) the $M$--transform serves better than the Green's function. However, it is customary to write the relationship between (\ref{eq:MeanSpectralDensityDefinition}) and (\ref{eq:GreenFunctionAndMTransformDefinition}) in terms of this latter, \begin{equation}\label{eq:SokhotskyFormula} \rho_{\mathbf{H}} ( \lambda ) = - \frac{1}{\pi} \lim_{\epsilon \to 0^{+}} \mathrm{Im} G_{\mathbf{H}} ( \lambda + \mathrm{i} \epsilon ) = - \frac{1}{2 \pi \mathrm{i}} \lim_{\epsilon \to 0^{+}} \left( G_{\mathbf{H}} ( \lambda + \mathrm{i} \epsilon ) - G_{\mathbf{H}} ( \lambda - \mathrm{i} \epsilon ) \right) . \end{equation} resulting from a well--known formula for generalized functions, \smash{$\lim_{\epsilon \to 0^{+}} 1 / ( x \pm \mathrm{i} \epsilon ) = \textrm{pv} ( 1 / x ) \mp \mathrm{i} \pi \delta ( x )$}. \subsubsection{The $N$--Transform and Free Random Variables} \label{sss:TheNTransformAndFreeRandomVariables} The doubly correlated Wishart ensemble $\mathbf{c}$ (\ref{eq:EstimatorcDefinition}) may be viewed as a product of several random and non--random matrices. The general problem of multiplying random matrices seems formidable. In classical probability theory, it can be effectively handled in the special situation when the random terms are independent: then, the exponential map reduces it to the addition problem of independent random numbers, solved by considering the logarithm of the characteristic functions of the respective PDFs, which proves to be additive. In matrix probability theory, a crucial insight came from D.~Voiculescu and coworkers and R.~Speicher~\cite{VoiculescuDykemaNica1992,Speicher1994}, who showed how to parallel the commutative construction in the noncommutative world. It starts with the notion of ``freeness,'' which basically comprises probabilistic independence together with a lack of any directional correlation between two random matrices. This nontrivial new property happens to be the right extension of classical independence, as it allows for an efficient algorithm of multiplying free random variables (FRV), which we state below: \begin{description} \item[Step 1:] Suppose we have two random matrices, \smash{$\mathbf{H}_{1}$} and \smash{$\mathbf{H}_{2}$}, mutually free. Their spectral properties are best wrought into the $M$--transforms (\ref{eq:GreenFunctionAndMTransformDefinition}), \smash{$M_{\mathbf{H}_{1}} ( z )$} and \smash{$M_{\mathbf{H}_{2}} ( z )$}. \item[Step 2:] The critical maneuver is to turn attention to the functional inverses of these $M$--transforms, the so--called ``$N$--transforms,'' \begin{equation}\label{eq:NTransformDefinition} M_{\mathbf{H}} \left( N_{\mathbf{H}} ( z ) \right) = N_{\mathbf{H}} \left( M_{\mathbf{H}} ( z ) \right) = z . \end{equation} \item[Step 3:] The $N$--transforms submit to a very straightforward rule upon multiplying free random matrices (the ``FRV multiplication law''), \begin{equation}\label{eq:FRVMultiplicationLaw} N_{\mathbf{H}_{1} \mathbf{H}_{2}} ( z ) = \frac{z}{1 + z} N_{\mathbf{H}_{1}} ( z ) N_{\mathbf{H}_{2}} ( z ) , \qquad \textrm{for free \smash{$\mathbf{H}_{1}$}, \smash{$\mathbf{H}_{2}$}.} \end{equation} \item[Step 4:] Finally, it remains to functionally invert the resulting $N$--transform \smash{$N_{\mathbf{H}_{1} \mathbf{H}_{2}} ( z )$} to gain the $M$--transform of the product, \smash{$M_{\mathbf{H}_{1} \mathbf{H}_{2}} ( z )$}, and consequently, all the spectral properties via formula~(\ref{eq:SokhotskyFormula}). \end{description} It is stunning that such a simple prescription (relying on the choice of the $M$--transform as the carrier of the mean spectral information, and the construction of its functional inverse, the $N$--transform, which essentially multiplies under taking the free product) resolves the multiplication problem for free random noncommutative objects. Let us just mention that the addition problem may be tackled along similar lines: In this case, the Green's function should be exploited, its functional inverse considered (\smash{$G_{\mathbf{H}} ( B_{\mathbf{H}} ( z ) ) = B_{\mathbf{H}} ( G_{\mathbf{H}} ( z ) ) = z$}; it is sometimes called the ``Blue's function''~\cite{Zee1996,JanikNowakPappZahed1997}), which obeys the ``FRV addition law,'' \smash{$B_{\mathbf{H}_{1} + \mathbf{H}_{2}} ( z ) = B_{\mathbf{H}_{1}} ( z ) + B_{\mathbf{H}_{2}} ( z ) - 1 / z$}, for two free random matrices. In this paper, we do not resort to using this addition formula, even though our problem could be approached through it as well. Let us also remark that in the original mathematical formulations~\cite{VoiculescuDykemaNica1992,Speicher1994} of these frames, a slightly different language is employed: Instead of the $N$--transform, the ``$S$--transform'' is found convenient, \smash{$S_{\mathbf{H}} ( z ) \equiv ( 1 + z ) / ( z N_{\mathbf{H}} ( z ) )$}, while in place of the Blue's function, one engages the ``$R$--transform,'' \smash{$R_{\mathbf{H}} ( z ) \equiv B_{\mathbf{H}} ( z ) - 1 / z$}. They fulfil simpler laws, \smash{$S_{\mathbf{H}_{1} \mathbf{H}_{2}} ( z ) = S_{\mathbf{H}_{1}} ( z ) S_{\mathbf{H}_{2}} ( z )$} and \smash{$R_{\mathbf{H}_{1} + \mathbf{R}_{2}} ( z ) = R_{\mathbf{H}_{1}} ( z ) + Y_{\mathbf{H}_{2}} ( z )$}, respectively. \subsubsection{Doubly Correlated Wishart Ensembles from Free Random Variables} \label{sss:DoublyCorrelatedWishartEnsemblesFromFreeRandomVariables} The innate potential of the FRV multiplication algorithm (\ref{eq:FRVMultiplicationLaw}) is surely revealed when inspecting the doubly correlated Wishart random matrix \smash{$\mathbf{c} = ( 1 / T ) \sqrt{\mathbf{C}} \widetilde{\mathbf{Y}} \mathbf{A} \widetilde{\mathbf{Y}}^{\mathrm{T}} \sqrt{\mathbf{C}}$} (\ref{eq:EstimatorcDefinition}). This has been done in detain in~\cite{BurdaJaroszJurkiewiczNowakPappZahed2009}, so we will only accentuate the main results here, referring the reader to the original paper for a thorough explanation. The idea is that one uses twice the cyclic property of the trace (which permits cyclic shifts in the order of the terms), and twice the FRV multiplication law (\ref{eq:FRVMultiplicationLaw}) (to break the $N$--transforms of products of matrices down to their constituents), in order to reduce the problem to solving the uncorrelated Wishart ensemble \smash{$( 1 / T ) \widetilde{\mathbf{Y}}^{\mathrm{T}} \widetilde{\mathbf{Y}}$}. This last model is further simplified, again by the cyclic property and the FRV multiplication rule applied once, to the standard $\mathbf{GOE}$ random matrix squared (and the projector \smash{$\mathbf{P} \equiv \mathrm{diag} ( \mathbf{1}_{N} , \mathbf{0}_{T - N} )$}, designed to chip the rectangle \smash{$\widetilde{\mathbf{Y}}$} off the square $\mathbf{GOE}$), whose properties are firmly established. Let us sketch the derivation, $$ N_{\mathbf{c}} ( z ) \stackrel{\substack{\mathrm{cyclic}\\\downarrow}}{=} N_{\frac{1}{T} \widetilde{\mathbf{Y}} \mathbf{A} \widetilde{\mathbf{Y}}^{\mathrm{T}} \mathbf{C}} ( z ) \stackrel{\substack{\mathrm{FRV}\\\downarrow}}{=} \frac{z}{1 + z} N_{\frac{1}{T} \widetilde{\mathbf{Y}} \mathbf{A} \widetilde{\mathbf{Y}}^{\mathrm{T}}} ( z ) N_{\mathbf{C}} ( z ) \stackrel{\substack{\mathrm{cyclic}\\\downarrow}}{=} $$ $$ \stackrel{\substack{\mathrm{cyclic}\\\downarrow}}{=} \frac{z}{1 + z} N_{\frac{1}{T} \widetilde{\mathbf{Y}}^{\mathrm{T}} \widetilde{\mathbf{Y}} \mathbf{A}} ( r z ) N_{\mathbf{C}} ( z ) \stackrel{\substack{\mathrm{FRV}\\\downarrow}}{=} \frac{z}{1 + z} \frac{r z}{1 + r z} N_{\frac{1}{T} \widetilde{\mathbf{Y}}^{\mathrm{T}} \widetilde{\mathbf{Y}}} ( r z ) N_{\mathbf{A}} ( r z ) N_{\mathbf{C}} ( z ) = $$ \begin{equation}\label{eq:DoublyCorrelatedWishartEnsembleTheMainEquation1} = r z N_{\mathbf{A}} ( r z ) N_{\mathbf{C}} ( z ) . \end{equation} This is the basic formula. Since the spectral properties of $\mathbf{c}$ are given by its $M$--transform, \smash{$M \equiv M_{\mathbf{c}} ( z )$}, it is more pedagogical to recast (\ref{eq:DoublyCorrelatedWishartEnsembleTheMainEquation1}) as an equation for the unknown $M$, \begin{equation}\label{eq:DoublyCorrelatedWishartEnsembleTheMainEquation2} z = r M N_{\mathbf{A}} ( r M ) N_{\mathbf{C}} ( M ) . \end{equation} It provides a means for computing the mean spectral density of a doubly correlated Wishart random matrix once the ``true'' covariance matrices $\mathbf{C}$ and $\mathbf{A}$ are given. In this communication, only a particular instance of this fundamental formula is applied, namely with an arbitrary auto--covariance matrix $\mathbf{A}$, but with trivial cross--covariances, \smash{$\mathbf{C} = \mathbf{1}_{N}$}. Using that \smash{$N_{\mathbf{1}_{K}} ( z ) = 1 + 1 / z$}, equation (\ref{eq:DoublyCorrelatedWishartEnsembleTheMainEquation2}) thins out to \begin{equation}\label{eq:DoublyCorrelatedWishartEnsembleTheMainEquation3} r M = M_{\mathbf{A}} \left( \frac{z}{r ( 1 + M )} \right) , \end{equation} which will be strongly exploited below. Let us mention that these equalities (\ref{eq:DoublyCorrelatedWishartEnsembleTheMainEquation2}), (\ref{eq:DoublyCorrelatedWishartEnsembleTheMainEquation3}) have been derived through other, more tedious, techniques (the planar Feynman--diagrammatic expansion, the replica trick) in~\cite{BurdaGorlichJaroszJurkiewicz2004,BurdaJurkiewicz2004,BurdaJurkiewiczWaclaw2005-1,BurdaGorlichJurkiewiczWaclaw2006,BurdaJurkiewiczWaclaw2005-2}. \section{VARMA from Free Random Variables} \label{s:VARMAFromFreeRandomVariables} In what follows, we will assume that the $\mathrm{VMA} ( q )$, $\mathrm{VAR} ( q )$, or $\mathrm{VARMA} ( q_{1} , q_{2} )$ stochastic processes are covariance (weak) stationary; for details, we refer to~\cite{Lutkepohl2005}. It implies certain restrictions on their parameters, but we will not bother with this issue in the current work. Another consequence is that the processes display some interesting features, such as invertibility. For all this, we must in particular take both $N$ and $T$ large from the start, with their ratio $r \equiv N / T$ fixed (\ref{eq:ThermodynamicalLimit}). More precisely, we stretch the range of the $a$--index from minus to plus infinity. This means that all the finite--size effects (appearing at the ends of the time series) are readily disregarded. In particular, there is no need to care about initial conditions for the processes, and all the recurrence relations are assumed to continue to the infinite past. \subsection{The $\mathrm{VMA} ( q )$ Process} \label{ss:TheVMAqProcess} \subsubsection{The Definition of $\mathrm{VMA} ( q )$} \label{sss:VMAqTheDefinition} We consider a situation when $N$ stochastic variables evolve according to identical independent $\mathrm{VMA} ( q )$ (vector moving average) processes, which we sample over a time span of $T$ moments. This is a simple generalization of the standard univariate weak--stationary moving average $\mathrm{MA} ( q )$. In such a setting, the value \smash{$Y_{i a}$} of the $i$--th ($i = 1 , \ldots , N$) random variable at time moment $a$ ($a = 1 , \ldots , T$) can be expressed as \begin{equation}\label{eq:VMAqDefinition} Y_{i a} = \sum_{\alpha = 0}^{q} a_{\alpha} \epsilon_{i , a - \alpha} . \end{equation} Here all the \smash{$\epsilon_{i a}$}'s are IID standard (mean zero, variance one) Gaussian random numbers (white noise), \smash{$\langle \epsilon_{i a} \epsilon_{j b} \rangle = \delta_{i j} \delta_{a b}$}. The \smash{$a_{\alpha}$}'s are some $( q + 1 )$ real constants; importantly, they do not depend on the index $i$, which reflects the fact that the processes are identical and independent (no ``spatial'' covariances among the variables). The rank $q$ of the process is a positive integer. \subsubsection{The Auto--Covariance Matrix} \label{sss:VMAqTheAutoCovarianceMatrix} In order to handle such a process (\ref{eq:VMAqDefinition}), notice that the \smash{$Y_{i a}$}'s, being linear combinations of uncorrelated Gaussian numbers, must also be Gaussian random variables, albeit displaying some correlations. Therefore, to fully characterize these variables, it is sufficient to calculate their two--point covariance function; this is straightforwardly done (see appendix~\ref{ss:TheAutoCovarianceMatrixForVMAq} for details), \begin{equation}\label{eq:VMAqTwoPointCovarianceFunction} \left\langle Y_{i a} Y_{j b} \right\rangle = \delta_{i j} A^{( 1 )}_{a b} , \end{equation} where \begin{equation}\label{eq:VMAqAutoCovarianceMatrix} A^{( 1 )}_{a b} = \kappa^{( 1 )}_{0} \delta_{a b} + \sum_{d = 1}^{q} \kappa^{( 1 )}_{d} \left( \delta_{a , b - d} + \delta_{a , b + d} \right) , \qquad \textrm{with} \qquad \kappa^{( 1 )}_{d} \equiv \sum_{\alpha = 0}^{q - d} a_{\alpha} a_{\alpha + d} , \qquad d = 0 , 1 , \ldots , q . \end{equation} In other words, the cross--covariance matrix is trivial, \smash{$\mathbf{C} = \mathbf{1}_{N}$} (no correlations between different variables), while the auto--covariance matrix \smash{$\mathbf{A}^{( 1 )}$}, responsible for temporal correlations, can be called ``$( 2 q + 1 )$--diagonal.'' In the course of this article, we will use several different auto--covariance matrices, and for brevity, we decide to label them with superscripts; their definitions are all collected in appendix~\ref{ss:AListOfTheVariousAutoCovarianceMatricesUsed}. For example, in the simplest case of $\mathrm{VMA} ( 1 )$, it is tri--diagonal, \begin{equation}\label{eq:VMAOneAutoCovarianceMatrix} A^{( 1 )}_{a b} = \left( a_{0}^{2} + a_{1}^{2} \right) \delta_{a b} + a_{0} a_{1} \left( \delta_{a , b - 1} + \delta_{a , b + 1} \right) . \end{equation} \subsubsection{The Fourier Transform and the $M$--Transform of the Auto--Covariance Matrix} \label{sss:VMAqTheFourierTransformAndTheMTransformOfTheAutoCovarianceMatrix} Such an infinite matrix (\ref{eq:VMAqAutoCovarianceMatrix}) is translationally invariant (as announced, it is one of the implications of the weak stationarity), \emph{i.e.}, the value of any of its entries depends only on the distance between its indices, \smash{$A^{( 1 )}_{a b} = A^{( 1 )} ( a - b )$}; specifically, \smash{$A^{( 1 )} ( \pm d ) = \kappa^{( 1 )}_{d}$}, for $d = 0 , 1 , \ldots , q$, and \smash{$A^{( 1 )} ( | d | > q ) = 0$}. Hence, it is convenient to rewrite this matrix in the Fourier space, \begin{equation}\label{eq:VMAqAutoCovarianceMatrixFourier} \widehat{A^{( 1 )}} ( p ) \equiv \sum_{d \in \mathbb{Z}} \mathrm{e}^{\mathrm{i} d p} A^{( 1 )} ( d ) = \kappa^{( 1 )}_{0} + 2 \sum_{d = 1}^{q} \kappa^{( 1 )}_{d} \cos ( d p ) . \end{equation} In this representation, the $M$--transform of \smash{$\mathbf{A}^{( 1 )}$} is readily obtained~\cite{BurdaJaroszJurkiewiczNowakPappZahed2009}, \begin{equation}\label{eq:MomentsGeneratingFunctionForATranslationallyInvariant} M_{\mathbf{A}^{( 1 )}} ( z ) = \frac{1}{2 \pi} \int_{- \pi}^{\pi} \mathrm{d} p \frac{\widehat{A^{( 1 )}} ( p )}{z - \widehat{A^{( 1 )}} ( p )} . \end{equation} This integral can be evaluated by the method of residues for any value of $q$, which we do in appendix~\ref{ss:TheMTransformOfTheAutoCovarianceMatrixForVMAq}, where also we print the general result (\ref{eq:VMAqGreenFunctionOfTheAutoCovarianceMatrix}). In particular, for $q = 1$, \begin{equation}\label{eq:VMAOneMomentsGeneratingFunctionForA} M_{\mathbf{A}^{( 1 )}} ( z ) = \frac{z}{\sqrt{z - \left( a_{0} + a_{1} \right)^{2}} \sqrt{z - \left( a_{0} - a_{1} \right)^{2}}} - 1 , \end{equation} where the square roots are principal. \subsubsection{The Pearson Estimator of the Covariances from Free Random Variables} \label{sss:VMAqThePearsonEstimatorOfTheCovariancesFromFreeRandomVariables} We will be interested in investigating the spectral properties of the Pearson estimator \smash{$\mathbf{c} = ( 1 / T ) \mathbf{Y} \mathbf{Y}^{\mathrm{T}} = ( 1 / T ) \widetilde{\mathbf{Y}} \mathbf{A}^{( 1 )} \widetilde{\mathbf{Y}}^{\mathrm{T}}$} (\ref{eq:EstimatorcDefinition}). The $M$--transform of this correlated Wishart random matrix, \smash{$M \equiv M_{\mathbf{c}} ( z )$}, can be retrieved from equation (\ref{eq:DoublyCorrelatedWishartEnsembleTheMainEquation3}). We could write it for any $q$ using (\ref{eq:VMAqGreenFunctionOfTheAutoCovarianceMatrix}), but we will restrict ourselves only to $q = 1$, in which case the substitution of (\ref{eq:VMAOneMomentsGeneratingFunctionForA}) leads to a fourth--order polynomial (Ferrari) equation for the unknown $M$, $$ r^{4} \left( a_{0}^{2} - a_{1}^{2} \right)^{2} M^{4} + 2 r^{3} \left( - \left( a_{0}^{2} + a_{1}^{2} \right) z + \left( a_{0}^{2} - a_{1}^{2} \right)^{2} (r + 1) \right) M^{3} + $$ $$ + r^{2} \left( z^{2} - 2 \left( a_{0}^{2} + a_{1}^{2} \right) (r + 2) z + \left( a_{0}^{2} - a_{1}^{2} \right)^{2} \left( r^{2} + 4 r + 1 \right) \right) M^{2} + $$ \begin{equation}\label{eq:VMAOneMainEquation} + 2 r \left( z^{2} - \left( a_{0}^{2} + a_{1}^{2} \right) (2 r + 1) z + \left( a_{0}^{2} - a_{1}^{2} \right)^{2} r (r + 1) \right) M + r \left( - 2 \left( a_{0}^{2} + a_{1}^{2} \right) z + \left( a_{0}^{2} - a_{1}^{2} \right)^{2} r \right) = 0 . \end{equation} The FRV technique allowed us therefore to find this equation in a matter of a few lines of a simple algebraic computation. It has already been derived in~\cite{JinWangMiaoHuang2009}, and (\ref{eq:VMAOneMainEquation}) may be verified to coincide with the version given in that paper. In~\cite{JinWangMiaoHuang2009}, the pertinent equation is printed before (A.6), and to compare the two, one needs to change their variables into ours according to $y \to 1 / r$, $x \to z / r$, and \smash{$\underline{m} \to - r ( 1 + M ) / z$}. The last equality means that \smash{$\underline{m}$} and $m$ of~\cite{JinWangMiaoHuang2009} correspond in our language to the Green's functions \smash{$- r G_{\mathbf{c}} ( z )$} and \smash{$- G_{\mathbf{a}} ( z / r )$}, respectively, where \smash{$\mathbf{a} = ( 1 / N ) \mathbf{Y}^{\mathrm{T}} \mathbf{Y}$} is the Pearson estimator dual to $\mathbf{c}$. As mentioned, a quick extension to the case of arbitrary $q$ is possible, however the resulting equations for $M$ will be significantly more complicated; for instance, for $q = 2$, a lengthy ninth--order polynomial equation is discovered. \subsection{The $\mathrm{VAR} ( q )$ Process} \label{ss:TheVARqProcess} \subsubsection{The Definition of $\mathrm{VAR} ( q )$} \label{sss:VARqTheDefinition} A set--up of $N$ identical and independent $\mathrm{VAR} ( q )$ (vector auto--regressive) processes is somewhat akin to (\ref{eq:VMAqDefinition}), \emph{i.e.}, we consider $N$ decoupled copies of a standard univariate $\mathrm{AR} ( q )$ process, \begin{equation}\label{eq:VARqDefinition} Y_{i a} - \sum_{\beta = 1}^{q} b_{\beta} Y_{i , a - \beta} = a_{0} \epsilon_{i a} . \end{equation} It is again described by the demeaned and standardized Gaussian white noise \smash{$\epsilon_{i a}$} (which triggers the stochastic evolution), as well as $( q + 1 )$ real constants \smash{$a_{0}$}, \smash{$b_{\beta}$}, with $\beta = 1 , \ldots , q$. As announced before, the time stretches to the past infinity, so no initial condition is necessary. Although at first sight (\ref{eq:VARqDefinition}) may appear to be a more involved recurrence relation for the \smash{$Y_{i a}$}'s, it is actually easily reduced to the $\mathrm{VMA} ( q )$ case: It remains to remark that if one exchanges the \smash{$Y_{i a}$}'s with the \smash{$\epsilon_{i a}$}'s, one precisely arrives at the $\mathrm{VMA} ( q )$ process with the constants \smash{$a^{( 2 )}_{0} \equiv 1 / a_{0}$}, \smash{$a^{( 2 )}_{\beta} \equiv - b_{\beta} / a_{0}$}, $\beta = 1 , \ldots , q$. In other words, the auto--covariance matrix \smash{$\mathbf{A}^{( 3 )}$} of the $\mathrm{VAR} ( q )$ process (\ref{eq:VARqDefinition}) is simply the inverse of the auto--covariance matrix \smash{$\mathbf{A}^{( 2 )}$} of the corresponding $\mathrm{VMA} ( q )$ process with the described modification of the parameters, \begin{equation}\label{eq:VARqFromVMAq} \mathbf{A}^{( 3 )} = \left( \mathbf{A}^{( 2 )} \right)^{- 1} . \end{equation} This inverse exists thanks to the weak stationarity supposition. \subsubsection{The Fourier Transform and the $M$--Transform of the Auto--Covariance Matrix} \label{sss:VARqTheFourierTransformAndTheMTransformOfTheAutoCovarianceMatrix} The Fourier transform of the auto--covariance matrix \smash{$\mathbf{A}^{( 3 )}$} of $\mathrm{VAR} ( q )$ is therefore a (number) inverse of its counterpart for $\mathrm{VMA} ( q )$ with its parameters appropriately changed, \begin{equation}\label{eq:VARqAutoCovarianceMatrixFourier} \widehat{A^{( 3 )}} ( p ) = \frac{1}{\widehat{A^{( 2 )}} ( p )} = \frac{1}{\kappa^{( 2 )}_{0} + 2 \sum_{d = 1}^{q} \kappa^{( 2 )}_{d} \cos ( d p )} , \end{equation} where \begin{equation}\label{eq:VARqKappas} \kappa^{( 2 )}_{d} = \frac{1}{a_{0}^{2}} \sum_{\alpha = 0}^{q - d} b_{\alpha} b_{\alpha + d} , \qquad d = 0 , 1 , \ldots , q , \end{equation} and where we define \smash{$b_{0} \equiv - 1$}. In order to find the $M$--transform of the inverse matrix, \smash{$\mathbf{A}^{( 3 )} = ( \mathbf{A}^{( 2 )} )^{- 1}$}, one employs a general result, true for any (real symmetric) random matrix $\mathbf{H}$, and obtainable through an easy algebra, \begin{equation}\label{eq:MHInverseFromMH} M_{\mathbf{H}^{- 1}} ( z ) = - M_{\mathbf{H}} ( 1 / z ) - 1 . \end{equation} Since the quantity \smash{$M_{\mathbf{A}^{( 2 )}} ( z )$} is known for any $q$ (\ref{eq:VMAqGreenFunctionOfTheAutoCovarianceMatrix}), hence is \smash{$M_{\mathbf{A}^{( 3 )}} ( z )$} via (\ref{eq:MHInverseFromMH}), but we will not print it explicitly. Let us just give it for $q = 1$, in which case (\ref{eq:MHInverseFromMH}) and (\ref{eq:VMAOneMomentsGeneratingFunctionForA}) yield \begin{equation}\label{eq:VAROneMomentsGeneratingFunctionForA} M_{\mathbf{A}^{( 3 )}} ( z ) = - \frac{1}{\sqrt{1 - \frac{\left( 1 - b_{1} \right)^{2}}{a_{0}^{2}} z} \sqrt{1 - \frac{\left( 1 + b_{1} \right)^{2}}{a_{0}^{2}} z}} . \end{equation} \subsubsection{The Auto--Covariance Matrix} \label{sss:VARqTheAutoCovarianceMatrix} Despite being somewhat outside of the main line of thought of this article, an interesting question would be to search for an explicit expression for the auto--covariance matrix \smash{$\mathbf{A}^{( 3 )}$} from its Fourier transform (\ref{eq:VARqAutoCovarianceMatrixFourier}), \begin{equation}\label{eq:VARqAutoCovarianceMatrixFromItsFourierTransform} A^{( 3 )} ( d ) = \frac{1}{2 \pi} \int_{- \pi}^{\pi} \mathrm{d} p \mathrm{e}^{- \mathrm{i} d p} \frac{1}{\kappa^{( 2 )}_{0} + 2 \sum_{l = 1}^{q} \kappa^{( 2 )}_{l} \cos ( l p )} , \end{equation} where we exploited the fact that \smash{$\mathbf{A}^{( 3 )}$} must be translationally invariant, \smash{$A^{( 3 )}_{a b} = A^{( 3 )} ( a - b )$}. This computation would shed light on the structure of temporal correlations present in a VAR setting. This integral is evaluated by the method of residues in a very similar manner to the one shown in appendix~\ref{ss:TheMTransformOfTheAutoCovarianceMatrixForVMAq}, and we do this in appendix~\ref{ss:TheAutoCovarianceMatrixForVARq}. We discover that the auto--covariance matrix is a sum of $q$ exponential decays, \begin{equation}\label{eq:VARqAutoCovarianceMatrix1} A^{( 3 )} ( d ) = \sum_{\gamma = 1}^{q} C_{\gamma} \mathrm{e}^{- | d | / T_{\gamma}} , \end{equation} where \smash{$C_{\gamma}$} are constants, and \smash{$T_{\gamma}$} are the characteristic times (\ref{eq:VARqAutoCovarianceMatrixCharacteristicTimes}), $\gamma = 1 , \ldots , q$; these constituents are given explicitly in (\ref{eq:VARqAutoCovarianceMatrix2}). This is a well--known fact, nevertheless we wanted to establish it again within our approach. For example, for $q = 1$, the auto--covariance matrix of $\mathrm{VAR} ( 1 )$ is one exponential decay, \begin{equation}\label{eq:VAROneAutoCovarianceMatrix} A^{( 3 )} ( d ) = \frac{a_{0}^{2}}{1 - b_{1}^{2}} b_{1}^{| d |} , \end{equation} where we assumed for simplicity \smash{$0 < b_{1} < 1$} (the formula can be easily extended to all values of \smash{$b_{1}$}). \subsubsection{The Pearson Estimator of the Covariances from Free Random Variables} \label{sss:VARqThePearsonEstimatorOfTheCovariancesFromFreeRandomVariables} Having found an expression for the $M$--transform of the auto--covariance matrix \smash{$\mathbf{A}^{( 3 )}$} of a $\mathrm{VAR} ( q )$ (\ref{eq:MHInverseFromMH}), (\ref{eq:VMAqGreenFunctionOfTheAutoCovarianceMatrix}), we may proceed to investigate the equation (\ref{eq:DoublyCorrelatedWishartEnsembleTheMainEquation3}) for the $M$--transform \smash{$M \equiv M_{\mathbf{c}} ( z )$} of the correlated Wishart random matrix \smash{$\mathbf{c} = ( 1 / T ) \mathbf{Y} \mathbf{Y}^{\mathrm{T}} = ( 1 / T ) \widetilde{\mathbf{Y}} \mathbf{A}^{( 3 )} \widetilde{\mathbf{Y}}^{\mathrm{T}}$} (\ref{eq:EstimatorcDefinition}). We will do this explicitly only for $q = 1$, when (\ref{eq:VAROneMomentsGeneratingFunctionForA}) leads to a fourth--order (Ferrari) polynomial equation for the unknown $M$, $$ a_{0}^{4} r^{2} M^{4} + 2 a_{0}^{2} r \left( - \left( 1 + b_{1}^{2} \right) z + a_{0}^{2} r \right) M^{3} + $$ \begin{equation}\label{eq:VAROneMainEquation} + \left( \left( 1 - b_{1}^{2} \right)^{2} z^{2} - 2 a_{0}^{2} r \left( 1 + b_{1}^{2} \right) z + \left( r^{2} - 1 \right) a_{0}^{4} \right) M^{2} - 2 a_{0}^{4} M - a_{0}^{4} = 0 . \end{equation} This equation has been derived by another method in~\cite{JinWangMiaoHuang2009}, and our result confirms their equation (A.8), with the change in notation, $y \to 1 / r$, $x \to z / r$, $z \to r M$. \subsection{The \smash{$\mathrm{VARMA} ( q_{1} , q_{2} )$} Process} \label{ss:TheVARMAq1q2Process} \subsubsection{The Definition of \smash{$\mathrm{VARMA} ( q_{1} , q_{2} )$}} \label{sss:VARMAq1q2TheDefinition} The two types of processes which we elaborated on above, \smash{$\mathrm{VAR} ( q_{1} )$} and \smash{$\mathrm{VMA} ( q_{2} )$}, can be combined into one stochastic process called \smash{$\mathrm{VARMA} ( q_{1} , q_{2} )$}, \begin{equation}\label{eq:VARMAq1q2Definition} Y_{i a} - \sum_{\beta = 1}^{q_{1}} b_{\beta} Y_{i , a - \beta} = \sum_{\alpha = 0}^{q_{2}} a_{\alpha} \epsilon_{i , a - \alpha} . \end{equation} Now it is a straightforward and well--known observation (which can be verified by a direct calculation) that the auto--covariance matrix \smash{$\mathbf{A}^{( 5 )}$} of this process is simply the product (in any order) of the auto--covariance matrices of the VAR and VMA pieces; more precisely, \begin{equation}\label{eq:VARMAq1q2FromVARq1VMAq2} \mathbf{A}^{( 5 )} = \left( \mathbf{A}^{( 4 )} \right)^{- 1} \mathbf{A}^{( 1 )} , \end{equation} where \smash{$\mathbf{A}^{( 1 )}$} corresponds to the generic \smash{$\mathrm{VMA} ( q_{2} )$} model (\ref{eq:VMAqAutoCovarianceMatrix}), while \smash{$\mathbf{A}^{( 4 )}$} denotes the auto--covariance matrix of \smash{$\mathrm{VMA} ( q_{1} )$} with a slightly different modification of the parameters compared to the previously used, namely \smash{$a^{( 4 )}_{0} \equiv 1$}, \smash{$a^{( 4 )}_{\beta} \equiv - b_{\beta}$}, for \smash{$\beta = 1 , \ldots , q_{1}$}. We have thus already made use here of the fact that the auto--covariance matrix of a VAR process is the inverse of the auto--covariance matrix of a certain corresponding VMA process (\ref{eq:VARqFromVMAq}), but the new change in parameters necessary in moving from VAR to VMA has effectively \smash{$a_{0} = 1$} w.r.t. what we had before (\ref{eq:VARqFromVMAq}); it is understandable: this ``missing'' \smash{$a_{0}$} is now included in the matrix of the other \smash{$\mathrm{VMA} ( q_{2} )$} process. \begin{figure}[t] \includegraphics[width=8.5cm]{VARMA11Varyingr.eps} \includegraphics[width=8.5cm]{VARMA11Varyinga1b1.eps} \includegraphics[width=8.5cm]{VARMA11Varyinga1.eps} \includegraphics[width=8.5cm]{VARMA11Varyingb1.eps} \caption{The mean spectral density \smash{$\rho_{\mathbf{c}} ( \lambda )$} of the Pearson estimator $\mathbf{c}$ of the cross--covariances in the $\mathrm{VARMA} ( 1 , 1 )$ process computed numerically from the sixth--order polynomial equation (\ref{eq:VARMAOneOneMainEquation}), for various values of the process' parameters. The scale of these parameters is determined by choosing \smash{$a_{0} = 1$} everywhere. Recall that the theoretical formula (\ref{eq:VARMAOneOneMainEquation}) is valid in the thermodynamical limit (\ref{eq:ThermodynamicalLimit}) of $N , T \to \infty$, with $r = N / T$ kept finite.\\UP LEFT: We set the remaining VARMA parameters to \smash{$a_{1} = 0.3$}, \smash{$b_{1} = 0.2$}, while the rectangularity ratio takes the values $r = 0.5$ (the purple line), $0.1$ (red), $0.02$ (magenta), $0.004$ (pink); each one is $5$ times smaller than the preceding one. We observe how the graphs become increasingly peaked (narrower and taller) around $\lambda = 1$ as $r$ decreases, which reflects the movement of the estimator $\mathbf{c}$ toward its underlying value \smash{$\mathbf{C} = \mathbf{1}_{N}$}.\\UP RIGHT: We fix $r = 0.25$ and consider the two VARMA parameters equal to each other, with the values \smash{$a_{1} = b_{1} = 0.6$} (purple), $0.4$ (red), $0.2$ (magenta), $0.01$ (pink).\\DOWN LEFT: We hold $r = 0.25$ and \smash{$b_{1} = 0.2$}, and modify \smash{$a_{1} = 0.6$} (purple), $0.4$ (red), $0.2$ (magenta), $0.0$ (pink); for this last value, the $\mathrm{VARMA} ( 1 , 1 )$ model reduces to $\mathrm{VAR} ( 1 )$.\\DOWN RIGHT: Similarly, but this time we assign $r = 0.25$ and \smash{$a_{1} = 0.2$}, while changing \smash{$b_{1} = 0.6$} (purple), $0.4$ (red), $0.2$ (magenta), $0.0$ (pink); this last value corresponds to $\mathrm{VMA} ( 1 )$.} \label{fig:VARMAOneOneTheory} \end{figure} \begin{figure}[h] \includegraphics[width=8.5cm]{VARMA11MC1000.eps} \includegraphics[width=8.5cm]{VARMA11MC10000.eps} \caption{Monte Carlo simulations of the mean spectral density \smash{$\rho_{\mathbf{c}} ( \lambda )$} (the green plots) compared to the theoretical result obtained numerically from the sixth--order equation (\ref{eq:VARMAOneOneMainEquation}) (the dashed red lines). The conformity is nearly perfect. We generate the matrices $\mathbf{Y}$ of sizes $N = 50$, $T = 200$ (\emph{i.e.}, $r = 0.25$) from the $\mathrm{VARMA} ( 1 , 1 )$ process with the parameters \smash{$a_{0} = 1$}, \smash{$a_{1} = 0.3$}, \smash{$b_{1} = 0.2$}. The Monte Carlo simulation is repeated $1,000$ (LEFT) or $10,000$ (RIGHT) times; in this latter case, a significant improvement in the quality of the agreement is seen. One notices finite--size effects at the edges of the spectrum (``leaking out'' of eigenvalues): in the numerical simulations, $N$ and $T$ are obviously finite, while equation (\ref{eq:VARMAOneOneMainEquation}) is legitimate in the thermodynamical limit (\ref{eq:ThermodynamicalLimit}) only, hence the small discrepancies; by enlarging the chosen dimensions $50 \times 200$ one would diminish this fallout.} \label{fig:VARMAOneOneTheoryPlusMC} \end{figure} \subsubsection{The Fourier Transform and the $M$--Transform of the Auto--Covariance Matrix} \label{sss:VARMAq1q2TheFourierTransformAndTheMTransformOfTheAutoCovarianceMatrix} The Fourier transform of the auto--covariance matrix \smash{$\mathbf{A}^{( 5 )}$} of \smash{$\mathrm{VARMA} ( q_{1} , q_{2} )$} (\ref{eq:VARMAq1q2FromVARq1VMAq2}) is simply the product of the respective Fourier transforms (\ref{eq:VMAqAutoCovarianceMatrixFourier}) and (\ref{eq:VARqAutoCovarianceMatrixFourier}), \begin{equation}\label{eq:VARMAq1q2AutoCovarianceMatrixFourier} \widehat{A^{( 5 )}} ( p ) = \frac{\kappa^{( 1 )}_{0} + 2 \sum_{d_{2} = 1}^{q_{2}} \kappa^{( 1 )}_{d_{2}} \cos \left( d_{2} p \right)}{\kappa^{( 4 )}_{0} + 2 \sum_{d_{1} = 1}^{q_{1}} \kappa^{( 4 )}_{d_{1}} \cos \left( d_{1} p \right)} , \end{equation} where \begin{equation}\label{eq:VARMAq1q2KappasAndLambdas} \kappa^{( 4 )}_{d_{1}} = \sum_{\alpha_{1} = 0}^{q_{1} - d_{1}} b_{\alpha_{1}} b_{\alpha_{1} + d_{1}} , \qquad \kappa^{( 1 )}_{d_{2}} = \sum_{\alpha_{2} = 0}^{q_{2} - d_{2}} a_{\alpha_{2}} a_{\alpha_{2} + d_{2}} , \qquad d_{1} = 0 , 1 , \ldots , q_{1} , \qquad d_{2} = 0 , 1 , \ldots , q_{2} , \end{equation} where we recall \smash{$b_{0} = - 1$}. For instance, for $\mathrm{VARMA} ( 1 , 1 )$ (it is described by three constants, \smash{$a_{0}$}, \smash{$a_{1}$}, \smash{$b_{1}$}), one explicitly has \begin{equation}\label{eq:VARMAOneOneAutoCovarianceMatrixFourier} \widehat{A^{( 5 )}} ( p ) = \frac{a_{0}^{2} + a_{1}^{2} + 2 a_{0} a_{1} \cos p}{1 + b_{1}^{2} - 2 b_{1} \cos p} . \end{equation} The $M$--transform of \smash{$\mathbf{A}^{( 5 )}$} can consequently be derived from the general formula (\ref{eq:MomentsGeneratingFunctionForATranslationallyInvariant}). We will evaluate here the pertinent integral only for the simplest $\mathrm{VARMA} ( 1 , 1 )$ process, even though an arbitrary case may be handled by the technique of residues, \begin{equation}\label{eq:VARMAOneOneMomentsGeneratingFunctionForAFive} M_{\mathbf{A}^{( 5 )}} ( z ) = \frac{1}{a_{0} a_{1} + b_{1} z} \left( - a_{0} a_{1} + \frac{z \left( a_{0} a_{1} + \left( a_{0}^{2} + a_{1}^{2} \right) b_{1} + a_{0} a_{1} b_{1}^{2} \right)}{\sqrt{\left( 1 - b_{1} \right)^{2} z - \left( a_{0} + a_{1} \right)^{2}} \sqrt{\left( 1 + b_{1} \right)^{2} z - \left( a_{0} - a_{1} \right)^{2}}} \right) . \end{equation} \subsubsection{The Auto--Covariance Matrix} \label{sss:VARMAq1q2TheAutoCovarianceMatrix} One might again attempt to track the structure of temporal covariances in a VARMA process. This can be done either by the inverse Fourier transform of (\ref{eq:VARMAq1q2AutoCovarianceMatrixFourier}), or through a direct computation based on the recurrence relation (\ref{eq:VARMAq1q2Definition}) (importantly, adhering to the assumption that it stretches to the past infinity). Let us print the result just for $\mathrm{VARMA} ( 1 , 1 )$, \begin{equation}\label{eq:VARMAOneOneAutoCovarianceMatrix} A^{( 5 )} ( d ) = - \frac{a_{0} a_{1}}{b_{1}} \delta_{d , 0} + \frac{\left( a_{1} + a_{0} b_{1} \right) \left( a_{0} + a_{1} b_{1} \right)}{b_{1} \left( 1 - b_{1}^{2} \right)} b_{1}^{| d |} , \end{equation} where for simplicity \smash{$0 < b_{1} < 1$}. This is an exponential decay, with the characteristic time of the VAR piece, with an additional term on the diagonal. \subsubsection{The Pearson Estimator of the Covariances from Free Random Variables} \label{sss:VARMAq1q2ThePearsonEstimatorOfTheCovariancesFromFreeRandomVariables} Expression (\ref{eq:VARMAOneOneMomentsGeneratingFunctionForAFive}), along with the fundamental FRV formula (\ref{eq:DoublyCorrelatedWishartEnsembleTheMainEquation3}), allow us to write the equation satisfied by the $M$--transform \smash{$M \equiv M_{\mathbf{c}} ( z )$} of the Pearson estimator \smash{$\mathbf{c} = ( 1 / T ) \mathbf{Y} \mathbf{Y}^{\mathrm{T}} = ( 1 / T ) \widetilde{\mathbf{Y}} \mathbf{A}^{( 5 )} \widetilde{\mathbf{Y}}^{\mathrm{T}}$} (\ref{eq:EstimatorcDefinition}) of the cross--covariances in the $\mathrm{VARMA} ( 1 , 1 )$ process; it happens to be polynomial of order six, and we print it (\ref{eq:VARMAOneOneMainEquation}) in appendix~\ref{ss:TheEquationForTheMTransformOfThePearsonEstimatorOfTheCovariancesForVARMAOneOne}. It may be solved numerically, a proper solution chosen (the one which leads to a sensible density: real, positive--definite, normalized to unity), and finally, the mean spectral density \smash{$\rho_{\mathbf{c}} ( \lambda )$} derived from (\ref{eq:SokhotskyFormula}). We show the shapes of this density for a variety of the values of the parameters $r$, \smash{$a_{0}$}, \smash{$a_{1}$}, \smash{$b_{1}$} in fig.~\ref{fig:VARMAOneOneTheory}. Moreover, in order to test the result (\ref{eq:VARMAOneOneMainEquation}), and more broadly, to further establish our FRV framework in the guise of formula (\ref{eq:DoublyCorrelatedWishartEnsembleTheMainEquation3}), the theoretical form of the density is compared to Monte Carlo simulations in fig.~\ref{fig:VARMAOneOneTheoryPlusMC}; they remain in excellent concord. These are the main findings of this article. \section{Conclusions} \label{s:Conclusions} In this paper we attempted to advertise the power and flexibility of the Free Random Variables calculus for multivariate stochastic processes of the VARMA type. The FRV calculus is ideally suited for multidimensional time series problems, provided the dimensions of the underlying matrices are large. The operational procedures are simple, algebraic and transparent. The structure of the final formula which relates the moments' generating function of the population covariance and the sample covariance allows one to easily derive eigenvalue density of the sample covariance. We in detail illustrated how this procedure works for $\mathrm{VARMA} ( 1 , 1 )$, confronted the theoretical prediction with numerical data obtained by Monte Carlo simulations of the VARMA process and observed a perfect agreement. The FRV calculus is not restricted to Gaussian variables. It also works for non--Gaussian processes, including those with heavy--tailed increments belonging to the L\'{e}vy basin of attraction, where the moments do not exist. Since the majority of data collected nowadays is naturally stored in the form of huge matrices, we believe that the FRV technique is the most natural candidate for the matrix--valued ``probability calculus'' that can provide efficient algorithms for cleaning (de--noising) large sets of data and unraveling essential but hidden correlations. \begin{acknowledgements} This work has been supported by the Polish Ministry of Science Grant No.~N~N202~229137 (2009--2012). AJ acknowledges the support of Clico Ltd. \end{acknowledgements}
\section{Introduction and Main Results} The rapid development of the theory of entanglement over the last decade or so \cite{HHHH-RMP}, and its usefulness in communication systems and computational devices, as well as the experimental observations of entangled states in a variety of distinct physical systems \cite{ref-boi}, have attracted a lot of attention from different branches of physics, including condensed matter and ultra-cold gases \cite{ref-reviews1, ref-reviews2}. It has been argued that entanglement can be used as a ``universal detector'' of quantum phase transitions, with most of the studies being on the behavior of \emph{bipartite} entanglement \cite{Wootters, logneg}. A more natural way to study the many-body systems would be to consider multipartite entanglement, as almost all naturally occurring multisite quantum states are genuinely multi-party entangled. Such an enterprise is however limited by the intricate nature of entanglement theory in the multisite scenario. In particular, only a few multisite entanglement measures are known, and moreover their computation are difficult \cite{HHHH-RMP}. Multipartite states can have different hierarchies according to their entanglement quality and quantity. The simplest example is for three-particle states, where there are fully separable, biseparable, and genuine multipartite entangled states. A measure of genuine multiparty entanglement, quantifies, so to say, the ``purest'' form multiparty entanglement. In this paper, we define an entanglement measure, called generalized geometric measure (GGM), that can detect and quantify \emph{genuine} multiparticle entanglement. Interestingly, the measure is computable for arbitrary pure states of multiparty systems in arbitrary dimensions and arbitrary number of parties, and therefore can turn out to be a useful tool to detect quantum many-body phenomena, like quantum phase transitions. In this respect, GGM has the potential of gaining the same status in applications of multiparty entanglement theory, as that of logarithmic negativity \cite{logneg} in the bipartite domain. As an initial testing ground, we use the GGM to successfully detect quantum phase transitions in the anistropic XY model on a chain of spin-1/2 particles \cite{Barouch-McCoy}. Our main aim however is to apply the measure to states of frustrated spin systems, for which the phase diagrams are not exactly known. Frustrated many-body systems are a center of interest in condensed matter physics due to the typically rich and novel phase diagrams in such systems. Moreover, experimental realizations of many metal oxides, including those exhibiting high-Tc superconductivity, typically have frustrated interactions in their Hamiltonians \cite{snajh-er-jor,5mlParacetamol-diyechhi}. As paradigmatic representatives of such systems, we consider (i) the quasi 2D antiferromagnetic \(J_1-J_2\) Heisenberg model with nearest neighbor couplings, \(J_1\), and next-nearest neighbor couplings, \(J_2\) \cite{Majumdar-Ghosh, White-Affleck, Mikeska}, and (ii) the frustrated \(J_1-J_2\) model on a square lattice \cite{snajh-er-jor} (see Fig. 1). \begin{figure}[h!] \label{fig-chhobi} \begin{center} \epsfig{figure= 2D_j1_j2.eps, height=.15\textheight,width=0.47\textwidth} \caption{(Color online.) Two-dimensional \(J_1-J_2\) model, with vertical and horizontal couplings, \(J_1\), and diagonal couplings, \(J_2\) The predicted phase diagram is also schematically shown.} \end{center} \end{figure} For studying such systems, we introduce an order parameter which is the difference between the GGM (\({\cal E}\)) and its second derivative with respect to the system parameter, \(\mu\), that drives the transitions in the system. We call the quantity as ``bound GGM'', and is given by \begin{equation} {\cal E}_B \equiv {\cal E} - \frac{d^2 {\cal E}}{d\mu^2}. \nonumber \end{equation} The ground state manifold of the quasi 2D \(J_1-J_2\) system is not known exactly, except at the Majumdar-Ghosh point \cite{Majumdar-Ghosh}, i.e. for \(\alpha = J_2/J_1 = 0.5 \), where the system is highly frustrated, and presents two dimer states as its ground states. However, exact diagonalization and group theoretical studies show that the system is gapless, and hence critical, in the weakly frustrated regime, namely \(0 \leq \alpha \lesssim 0.24\) \cite{Majumdar-Ghosh, White-Affleck}. For higher coupling ratio \(\alpha\), the system enters a dimerized regime, and is gapped \cite{Majumdar-Ghosh, eita-AKLT}. We study the GGM for this system by exact diagonalization, and show that the bound GGM vanishes at the fluid-dimer transition point \(\alpha \approx 0.24\). Note here that it is known that bipartite entanglement cannot detect the gapless phase \cite{bipartite_MG} (cf. \cite{Chhajlany}). We find that the bound GGM is positive in the gapless phase while it becomes negative in the gapped one. The Majumdar-Ghosh point can also be detected by the GGM. Finally we apply our measure of genuine multisite entanglement to the ground state of the 2D Heisenberg system. As depicted in Fig. 1, the N{\' e}el and collinear ordered phases (the gapless phases dissociated by a phase having a finite gap between the singlet ground state and the excited states. We show that the bound GGM can detect both the quantum phase transitions -- from the N{\' e}el phase to the dimerized one at \(\alpha \approx 0.38\), as well as the transition from the dimer to the collinear phase at \(\alpha \approx 0.69\), as predicted, even for relatively small system-size. Like in the \(J_1-J_2\) ring, the positivity (negativity) of the bound GGM indicates the gapless (gapped) phase. Armed with these findings, we propose that the bound GGM can potentially be used for detecting gapped/gapless phases in many-body systems: \begin{eqnarray} \label{onek-deri-hoye-gyalo} {\cal E}_B > 0 \Rightarrow \mbox{gapless}, \quad {\cal E}_B < 0 \Rightarrow \mbox{gapped}. \end{eqnarray} This leads to an analogy with the thermodynamics of bound entanglement \cite{wwwww, HHHH-RMP}. Analogous to the first law of thermodynamics, internal energy = free energy + work done, a thermodynamic equation of entanglement was written: Entanglement cost = distillable entanglement + bound entanglement, where the bound entanglement is the amount of entanglement necessary to keep the transition (under local quantum operations and classical communication) from becoming irreversible. As another face of this entanglement-energy analogy, a negative value of \({\cal E}_B\), assuming the thesis in Eq. (\ref{onek-deri-hoye-gyalo}), indicates that the system needs a nonzero amount of energy to free itself from its ground state. We hope that this can help us in a quantification of the first law of the emerging entanglement thermodynamics \cite{thermodynamics, HHHH-RMP}. This is the reason for calling \({\cal E}_B\) as \emph{bound} GGM \cite{qqqqq}. \section{Generalized Geometric Measure} Let us begin by defining the generalized geometric measure. As mentioned above, GGM will quantify the genuineness of multiparty entanglement. An \(N\)-party pure quantum state is said to be genuinely \(N\)-party entangled, if it is not a product across any bipartite partition. The simplest examples of genuine tripartite entangled states are the Greenberger-Horne-Zeilinger \cite{GHZ} and W \cite{W-state} states. The GGM of an \(N\)-party pure quantum state \(|\psi\rangle\) is defined as \begin{equation} {\cal E} ( |\psi\rangle ) = 1 - \Lambda^2_{max} (|\psi\rangle ), \end{equation} where \(\Lambda_{max} (|\psi\rangle ) = \max | \langle \phi|\psi\rangle |\), with the maximization being over all pure states \(|\phi\rangle\) that are not genuinely \(N\)-party entangled. Note that the maximization performed in GGM is different from the maximization in the geometric measure of Ref. \cite{GM} (cf. \cite{hierarchy}). \subsection{Properties} Clearly, \({\cal E}\) is vanishing for all pure multiparty states that are not genuine multiparty entangled, and non-vanishing for others. We considered this quantity for four-party states in Ref. \cite{amadertele}, and showed it to be a mono- tonically decreasing quantity under local quantum operations and classical communication (LOCC). Applications of GGM to quantum many-body systems requires us to find its properties for an arbitrary number of parties. Let \(|\psi\rangle\) be an \(N\)-party pure quantum state in the tensor product Hilbert space \({\cal H}_{A_1} \otimes {\cal H}_{A_2} \otimes \ldots \otimes {\cal H}_{A_N}\). Therefore, the maximization in \begin{equation}\Lambda_{\max}(|\psi\rangle_{A_1 A_2 \ldots A_N}) = \max_{|\phi\rangle_{A_1 A_2 \ldots A_N}} |\langle \phi | \psi \rangle| \end{equation} is over all pure quantum states \(|\phi\rangle_{A_1 A_2 \ldots A_N}\), in \({\cal H}_{A_1} \otimes {\cal H}_{A_2} \otimes \ldots \otimes {\cal H}_{A_N}\), that are not genuinely multiparty entangled, which is a rather large class of states. Note however, that the square of \(\Lambda_{\max}(|\psi\rangle_{A_1 A_2 \ldots A_N})\) can be interpreted as the Born probability of some outcome in a quantum measurement on the state \(|\psi\rangle\). Now, entangled measurements cannot be worse than the product ones for any set of subsystems. Therefore, in the maximization, we do not need to consider the \(|\phi\rangle_{A_1 A_2 \ldots A_N}\) that are product in a partition of \(A_1, A_2, \ldots, A_N\) into three, four, ... sets. The only \(|\phi\rangle_{A_1 A_2 \ldots A_N}\) that are to be considered are the ones that are a product in a \emph{bi}-partition of \(A_1, A_2, \ldots, A_N\). This greatly reduces the class over which the maximization is carried out. Let \({\cal A}: {\cal B}\) be such a bi-partition. Then, \(\max |\langle \phi | \psi \rangle|\), where the maximization is carried over the \(|\phi\rangle\) that are product across \({\cal A}: {\cal B}\), is the maximal Schmidt coefficient, \(\lambda_{{\cal A}: {\cal B}}\), of the state \(|\psi\rangle_{A_1 A_2 \ldots A_N}\) in the \({\cal A}: {\cal B}\) bipartite split. \(\Lambda_{\max}(|\psi\rangle_{A_1 A_2 \ldots A_N})\) is therefore the maximum of all such maximal Schmidt coefficients in bipartite splits. Note that the \(\lambda\)'s involved in this closed form for \(\Lambda_{max}\) are all increasing under LOCC \cite{Vidal-Nielsen}. We have therefore proven the following theorem.\\ \noindent \textbf{Theorem.} \emph{The generalized geometric measure of \(|\psi\rangle_{A_1 A_2 \ldots A_N}\) is given by \begin{equation} {\cal E}(|\psi\rangle) = 1 - \max \{\lambda^2_{{\cal A}: {\cal B}} | {\cal A} \cup {\cal B} = \{1,2,\ldots, N\}, {\cal A} \cap {\cal B} = \emptyset\}. \end{equation} It is computable for a multiparty pure state of an arbitrary number of parties, and of arbitrary dimensions. Also, it is monotonically decreasing under LOCC. } \section{Anisotropic XY model} The one-dimensional XY model with \(N\) lattice sites is described by the Hamiltonian \begin{equation} \label{eq_XY_H} H_{XY} = \frac{J}{2} \left(\sum_{i=1}^{N} (1 + \gamma) \sigma^x_i \sigma^x_{i+1} + (1 - \gamma) \sigma^y_i \sigma^y_{i+1}\right) + h \sum_{i=1}^{N} \sigma_i^z, \end{equation} where \(J\) is the coupling constant, \(\gamma \in [0,1] \) is the anistropy parameter, \(\sigma\)'s are the Pauli matrices, and \(h\) represents the magnetic field in the transverse direction. The quantum transverse Ising and the transverse XX models correspond to two extreme values of \(\gamma\), which are resepectively \(\gamma =1\) and \(\gamma =0\). This model can be diagonalized by the Jordan-Wigner transformation \cite{Barouch-McCoy}. Apart from its other interests, it is the simplest model which shows a \emph{quantum} phase transition, driven by the magnetic field, at zero temperature. It is known to be detectable by using bipartite entanglement measures \cite{fazio-Nielsen}, like concurrence \cite{Wootters}. However, evaluating GGM will additionally quantify the nature of genuine multiparty entanglement of the ground state in this model, especially as it crosses the transition point. The diagonalization of this model can be achieved by introducing the Majorana fermions \begin{equation} c_{2l -1} = (\Pi_{i=1}^{l-1} \sigma_i^z)\sigma^x_l; \quad c_{2l} = (\Pi_{i=1}^{l-1} \sigma_i^z)\sigma^y_l. \end{equation} The Hamiltonian in Eq. (\ref{eq_XY_H}) thereby reduces to a quadratic fermionic Hamiltonian \cite{Barouch-McCoy}. The eigenvalues of the reduced density matrix of \(L\) sites of the ground state of this system can be obtained by using the above formalism \cite{ref-reviews2}, and is given by \begin{equation} e_{x_1 x_2 \ldots x_l} = \prod_{i=1}^{L} \frac{1 + (-1)^{x_i} \nu_i}{2}, \quad x_i = 0, 1 \phantom{,} \forall i, \end{equation} where \(\nu_i\)'s are the eigenvalues of \(G_L\), which in turn is given by \(B_L = G_L \otimes \left[\begin{array}{cc} 0 & 1 \\ -1 & 0 \\ \end{array}\right]\), with \begin{equation} \label{beRal-er-talobya-sho} G_L=\left[ \begin{array}{cccc} g_0 & \cdot & \cdot & g_{L-1} \\ \cdot & \cdot & \cdot & \cdot\\ - g_{L-1} & \cdot & \cdot & g_{0} \\ \end{array}\right], B_L = \left[ \begin{array}{cccc} \Pi_0 & \cdot & \cdot & \Pi_{L-1} \\ \cdot & \cdot & \cdot & \cdot\\ - \Pi_{L-1} & \cdot & \cdot & \Pi_{0} \\ \end{array}\right]. \nonumber \end{equation} Here, \(\Pi_l = \left[ \begin{array}{cc} 0 & g_l \\ -g_{-l} & 0 \\ \end{array}\right]\), and the real coefficients, \(g_l\), are given by \begin{equation} g_l = \frac{1}{2 \pi} \int_{0}^{2 \pi} d\phi e^{- i l \phi} \frac{\cos \phi - \lambda - i \gamma \sin \phi}{|\cos \phi - \lambda - i \gamma \sin \phi|}, \end{equation} where \(\lambda = J/h\). The derivative of GGM of the ground state, for different anistropy parameters \(\gamma\), clearly shows a logarithmic divergence at the transverse field given by \(\lambda =1\), as seen in Fig. 2. Note also that the ground state of the transverse Ising model (\(\gamma =1\)) has higher genuine multipartite entanglement as compared to the ground states for other values of \(\gamma\). This result may help us to understand the success of the dynamical states of the transverse Ising model as a substrate for efficient quantum computation \cite{raus-briegel}. \vspace{0.7cm} \begin{figure}[h!] \label{fig-chhobi-ek} \begin{center} \epsfig{figure= XY_GGM.eps, height=.27\textheight,width=0.45\textwidth} \caption{(Color online.) GGM of the transverse XY model. The GGM (actually \({\cal E} - 1/2\)) and its derivative (both dimensionless) are plotted on the vertical axis for the anistropic transverse XY model for different anisotropy parameters \(\gamma\), against the dimensionless system parameter \(\lambda\) on the horizontal axis. The dashed (blue), circled (pink), and dotted (green) lines are respectively for the Ising (\(\gamma =1\)), \(\gamma = 0.8\), and \(\gamma = 0.2\) models. The derivatives of GGM diverges at the quantum critical point \(\lambda = 1\). } \end{center} \end{figure} \section{Quasi 2D Frustrated \(J_1-J_2\) Model} We will now consider the frustrated quasi two-dimensional \(J_1-J_2\) Heisenberg model, in the case when both the nearest neighbor couplings, \(J_1\), and the next-nearest neighbor couplings, \(J_2\), are antiferromagnetic. Apart from its other interests, the intense interest for studying this model lies in the fact that it is similar to real systems, like \(\mbox{SrCuO}_{2}\) \cite{experiment}. The Hamiltonian of this model, with \(N\) lattice sites on a chain, is \begin{equation} H_{1D} = J_1 \sum_{i=1}^{N} \vec{\sigma}_i \cdot \vec{\sigma}_{i+1} + J_2 \sum_{i=1}^N \vec{\sigma}_i \cdot \vec{\sigma}_{i+2}, \end{equation} where \(J_1\) and \(J_2\) are both positive, and where periodic boundary condition in assumed. The ground state and the energy gap of this model were studied by using exact diagonalization, density matrix renormalization group method, bosonization technique, etc \cite{Mikeska}. For an even number of sites, the ground state at the Majumdar-Ghosh point (\(\alpha = J_2/J_1 = 0.5\)), is doubly degenerate, and the ground state manifold is spanned by the two dimers \(|\psi_{MG}^{\pm}\rangle = \Pi_{i=1}^{N/2} (|0\rangle_{2i} |1\rangle_{2i \pm 1} - |1\rangle_{2i} |0\rangle_{2i \pm 1})\), and the model is gapped at this point \cite{Majumdar-Ghosh}. For \(\alpha=0\), the Hamiltonian reduces to the \(s=1/2\) Heisenberg antiferromagnet and hence the the ground state, which is a spin fluid state having gapless excitations \cite{Griffiths-Yang}, can be obtained by Bethe ansatz. It is known that at \(\alpha \approx 0.2411\), a phase transtion from fluid to dimerization occurs \cite{gap_transition}. The genuine multipartite entanglement measure clearly signals the Majumdar-Ghosh point (See Fig. 3). The fluid-dimer tranition at \(\alpha \approx 0.24\) can also be detected by the vanishing of the bound GGM as the order parameter (for \(\mu = \alpha\)) (see Fig. 3). Moreover, \({\cal E}_B >0\) signals the gapless phase, while \({\cal E}_B <0\) indicates the gapped phase. \vspace{0.7cm} \begin{figure}[h!] \label{fig-chhobi-dui} \begin{center} \epsfig{figure= oneD_j1_j2_bound.eps, height=.27\textheight,width=0.45\textwidth} \caption{ (Color online.) GGM and bound GGM for the quasi-2D frustrated antiferromagnet. The left figure is for the GGM on the vertical axis against \(\alpha\) on the horizontal. The Majumdar-Ghosh point at \(\alpha = 0.5\) is clearly signaled. The figure on the right is for the bound GGM on the vertical axis, against \(\alpha\) on the horizontal, and the fluid-dimer transition is signaled by the vanishing of the bound GGM. The two curves are for \(8\) (red circles) and \(10\) (blue squares) spins in both the figures. The GGM, bound GGM, and \(\alpha\) are all dimensionless. } \end{center} \end{figure} \section{2D Frustrated \(J_1-J_2\) Model} We now consider spin-1/2 particles on a square lattice, where nearest neighbor spins (both vertical and horizontal) on the lattice are coupled by Heisenberg interactions, with coupling strengths \(J_1\), and where all diagonal spins are coupled by Heisenberg interactions, with coupling strengths \(J_2\) (see Fig. 1). This 2D model have attracted a lot of interest \cite{many_theory} due to its connection with the high \(T_c\)-superconductors and its similarity with magnetic materials like \(\mbox{Li}_2 \mbox{VOSiO}_4\) and \(\mbox{Li}_{2}\mbox{VOGeO}_4\) \cite{synthesis}. Although the different phases of the ground state of this model is well-studied, there seem to exist reasons to believe in further secrets hidden. The Hamiltonian of the system is therefore given by \begin{equation} H_{2D} = J_1 \sum_{\langle i,j \rangle} \vec{\sigma}_i \cdot \vec{\sigma}_{j} + J_2 \sum_{i,j \in {\cal D}} \vec{\sigma}_i \cdot \vec{\sigma}_{j}. \end{equation} Both \(J_1\) and \(J_2\) are antiferromagnetic (\(>0\)). In the classical limit, the model exhibits only a first-order phase transition from N{\' e}el to collinear at \(\alpha = J_2/J_1 = 0.5\). The phase diagram changes its nature, when quantum fluctuations are present, and in this case, the exact phase boundaries are not known. It is expected that two long range ordered (LRO) ground state phases are separated by quantum paramagnetic phases without LRO. Different methods, like exact diagonalization, series expansion methods, field-theory methods \cite{Richter10-Kim-Singh}, etc., applied to this model, predict that the first transition from N{\' e}el to dimer accurs at \(\alpha \approx 0.38\) while other one happens at \(\alpha \approx 0.66\). Recent experimental observations and proposals of detecting such phases in the laboratory demand the precise quantification of the phase diagram of this model at low temperature. Towards this aim, we show that even for relatively small system size, the order parameter based on the genuine multipartite entanglement measure (the \({\cal E}_B\), introduced above) can detect and quantify the phase diagram accurately. We perform exact diagonalization to find the ground state of the model, and we show that both the transitions -- N{\' e}el to dimer and dimer to collinear can be signaled by the bound GGM. A synopsis of these facts is given in Fig. 4. Precisely, we have found that \({\cal E}_B\) vanishes at \(\alpha \approx 0.38\) and again at \(\alpha \approx 0.69\). As observed in the case of the quasi 2D \(J_1-J_2\) model, \({\cal E}_B\) is positive in the gapless phases while it is negative in the intermediate gapped phase. \vspace{0.7cm} \begin{figure}[h!] \label{fig-chhobi-tin} \begin{center} \epsfig{figure= 2d_j1_j2_bound.eps, height=.27\textheight,width=0.45\textwidth} \caption{ (Color online.) GGM and bound GGM for the 2D frustrated antiferromagnet. The horizontal axes in both the figures represent \(\alpha\). The left figure is for the GGM on the vertical axis while the right one is for the bound GGM. The vanishing of the bound GGM signals both the N{\' e}el-to-dimer and the dimer-to-collinear transitions. And the gapped (gapless) phase(s) is (are) signaled by a negative (positive) bound GGM. The two curves are for \(9\) spins on a \(3 \times 3\) square lattice (red circles), and \(12\) spins on a \(3 \times 4\) rectangular lattice (blue squares) in both the figures. The GGM, bound GGM, and \(\alpha\) are all dimensionless. } \end{center} \end{figure} \section{Conclusions} Multipartite entangled states can be classified according to their separability in different partitions. Due to the complex classification, it is hard to obtain a unique multipartite entanglement measure. Instead of quantifying all the classes of multipartite states, we define an entanglement measure, called generlized geometric measure, that quantifies the ``purest'' form of multiparty entanglement, the genuine multipartite entanglement. This is akin to the situation in bipartite pure states, where there is essentially a unique entanglement measure, while mixed bipartite states allows a number of such measures \cite{HHHH-RMP}. In the case of multiparty states, we find ``pure'' and ``non-pure'' forms of entanglement, even within the class of pure states, where the ``pure'' part can be quantified by the generalized geometric measure defined here. Moreover, we found that the measure can be reduced to a simplified closed form, and hence is computable for arbitrary dimensions and arbitrary number of parties. We then applied this measure to detect phase diagrams in quantum many-body systems. After successfully verifying that the measure can detect quantum fluctuation driven phase transitions in the exactly solvable models like the XY Hamiltonian, we applied the generalized geometric measure to frustrated models like quasi two dimensional and two dimensional antiferromagnetic \(J_1-J_2\) models. In the latter case, the phase diagram is not known exactly, although there has been several predictions by different methods. In this paper, we show that an order parameter, called bound GGM, based on the multi-site entanglement measure defined, can signal the phase boundaries in both the models. Moreover, we found the the order parameter is positive when the system is gapless and negative in the gapped phase. We propose that the sign of the bound GGM can indicate whether a many-body system is gapped or gapless, and point to its implication for the first law of entanglement thermodynamics. \acknowledgments We acknowledge partial support from the Spanish MEC (TOQATA (FIS2008-00784)). \vspace{1cm}
\section{Introduction}\label{sec:intro} One of several basic open problems about Lindel\"{o}f spaces asks what possible cardinalities are for Lindel\"{o}f spaces in which each point is $\operatorname{\mathsf{G}}_{\delta}$. A number of consistency results using forcing or large cardinal axioms have been obtained. A fundamental issue that emerged from that work is the following question: \begin{quote} \emph{ When does a forcing extension preserve the Lindel\"{o}f property?} \end{quote} Surprisingly, little seems to be known about this question. Tall \cite{Tall:lindelofpointsgd} introduced the notion of \emph{indestructibly Lindel\"{o}f spaces}. A Lindel\"{o}f space is called an indestructibly Lindel\"{o}f space if it is still Lindel\"{o}f after forcing with any countably closed poset. Tall pointed out that a Lindel\"{o}f space is indestructibly Lindel\"{o}f if it remains Lindel\"{o}f after forcing with the poset which adjoins a Cohen subset of $\omega_1$ with countable conditions, which is just a particular instance of a countably closed poset. The class of spaces with the \emph{Rothberger property} is a natural and important subclass of the class of indestructibly Lindel\"{o}f spaces. It is of great interest to know which forcing notions preserve the Rothberger property. Scheepers and Tall showed \cite{ST:lindelof} that forcing with countably closed posets as well as the measure algebra preserve the Rothberger property. Both indestructible Lindel\"{o}f property and the Rothberger property are nicely characterized in terms of infinite games played on topological spaces. Pawlikowski \cite{Pa:pointopen} proved the Rothberger property is equivalent to the non-existence of a winning strategy for the first player in a certain game played on the space. Scheepers and Tall \cite{ST:lindelof} proved that an indestructibly Lindel\"{o}f space is characterized as a space on which the first player has no winning strategy in a modification of the game which appears in Pawlikowski's theorem into transfinite length. Moreover, the existence of a winning strategy for the second player in the game for indestructibly Lindel\"{o}f spaces also determines a noteworthy class of spaces, for it is known that if a space in which each point is $\operatorname{\mathsf{G}}_{\delta}$ belongs to this class then its cardinality is at most $2^{\aleph_0}$ \cite[Theorem~2]{ST:lindelof}. On the other hand, infinite games played on posets have been studied by many researchers, mainly in connection with Boolean-algebraic or forcing-theoretic properties. One of the most significant results of those studies is a game-theoretic characterization of proper forcing notions. Also the relations between game-theoretic properties and various properties of forcing notions, such as countable closedness, semiproperness, $\alpha$-properness, Axiom A, the Sacks property and the Laver property, have been studied by Foreman, Jech, Veli\v{c}kovi\'c, Zapletal, Shelah, Ishiu, Kada and others. See \cite{Fo:gameba,Ishiu:alphaproper,Je:moregame,JeSh:cccba,% Kada:gamecd,Vel:playful,Zap:morecg} for further information. In the present paper we show that, an indestructibly Lindel\"{o}f space remains Lindel\"{o}f after forcing with a poset in a natural class, which is described using a game and larger than the class of countably closed posets. Also we show that the Rothberger property is preserved under forcing with a poset in another natural class, which is again described using a game. We also investigate preservation of being a Lindel\"{o}f space in the class of P-spaces. Forcing with a proper poset preserves being a P-space. We will show that a Lindel\"{o}f P-space remains Lindel\"{o}f after forcing with a poset in a large class of proper posets. It is an intriguing question if there is an example of a Lindel\"{o}f P-space which is no longer a Lindel\"{o}f space after forcing with some proper poset (see Question~\ref{q:destroylindelofp}). In Section~\ref{sec:main} we establish a general preservation theorem stated in terms of games, which yields all the preservation results mentioned above. We will prove this theorem by pursuing moves in two games played in parallel, one is played on a topological space and the other on a poset. The general investigation of preservation of the Lindel\"{o}f property and its strengthenings under forcing extension has internal appeal, but the results may be useful in obtaining consistency results about a number of other basic open problems about Lindel\"{o}f spaces. \section{Preliminaries}\label{sec:pre} For a poset $\mathbb{P}$, an ordinal $\alpha$ and a cardinal $\kappa$, the \emph{cut-and-choose game $\operatorname{\mathsf{CG}}^{<\alpha}({<}\kappa)$ on $\mathbb{P}$} is defined as follows. The game is played by two players \textsc{One}\xspace and \textsc{Two}\xspace for $\alpha$ innings. In the beginning \textsc{One}\xspace chooses $p\in\mathbb{P}$. In each inning $\beta<\alpha$, \textsc{One}\xspace chooses a $\mathbb{P}$-name $\dot{\eta}_{\beta}$ for an ordinal, and then \textsc{Two}\xspace chooses a set $C_{\beta}$ of ordinals with $\size{C_{\beta}}<\kappa$. \textsc{Two}\xspace wins in this game if for every $\gamma<\alpha$ there is $q_{\gamma}\in\mathbb{P}$ such that $q_{\gamma}\leq p$ and $q_{\gamma}\forcestext[\mathbb{P}]{\forall \beta<\gamma\, (\dot{\eta}_{\beta}\in \check{C}_{\beta})}$. Note that, for \textsc{Two}\xspace to win, it is not required to find $q\in\mathbb{P}$ such that $q\leq p$ and $q\forcestext[\mathbb{P}]{\forall \beta<\alpha\, (\dot{\eta}_{\beta}\in \check{C}_{\beta})}$. Sometimes we preliminarily fix \textsc{One}\xspace's beginning move $p\in\mathbb{P}$ and then start the innings; in such a case we call it the \emph{game $\operatorname{\mathsf{CG}}^{<\alpha}({<}\kappa)$ on $\mathbb{P}$ below $p$}. If $\alpha=\gamma+1$, we write $\operatorname{\mathsf{CG}}^{\gamma}({<}\kappa)$ instead of $\operatorname{\mathsf{CG}}^{<\gamma+1}({<}\kappa)$. Also, we write $\operatorname{\mathsf{CG}}^{<\alpha}(\lambda)$ instead of $\operatorname{\mathsf{CG}}^{<\alpha}({<}\lambda^{+})$. The following theorem is well-known \cite{Je:moregame}. \begin{thm}\label{thm:countablechoiceproper} For a forcing notion $\mathbb{P}$, if \textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{CG}}^{\omega}(\aleph_0)$ on $\mathbb{P}$, then $\mathbb{P}$ is proper. \end{thm} We say a forcing notion $\mathbb{P}$ is \emph{$\omega^\omega$-bounding} if, for $p\in\mathbb{P}$ and a $\mathbb{P}$-name $\dot{f}$ for an element of $\omega^\omega$, there are $q\in\mathbb{P}$ and $g\in\omega^\omega$ such that $q\leq p$ and $q\forcestext[\mathbb{P}]{% \forall n<\omega\,(\dot{f}(n)\leq g(n))}$. The following fact is easily checked. \begin{thm}\label{thm:finitechoicebounding} For a forcing notion $\mathbb{P}$, if \textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{CG}}^{\omega}({<}\aleph_0)$ on $\mathbb{P}$, then $\mathbb{P}$ is $\omega^\omega$-bounding. \end{thm} \begin{rem} Although the converse of Theorem~\ref{thm:finitechoicebounding} does not hold, most well-known proper $\omega^\omega$-bounding forcing notions, such as Sacks forcing, Silver forcing and the measure algebra, are the ones on which \textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{CG}}^{\omega}({<}\aleph_0)$. \end{rem} A poset $\mathbb{P}$ is \emph{${<}\alpha$-closed} if any descending sequence in $\mathbb{P}$ of length less than $\alpha$ has a lower bound in $\mathbb{P}$. If $\mathbb{P}$ is ${<}\alpha$-closed, then obviously \textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{CG}}^{<\alpha}(1)$ on $\mathbb{P}$. \begin{rem} The game $\operatorname{\mathsf{CG}}^{<\alpha}(1)$ on a poset $\mathbb{P}$ is closely related to the \emph{strategic closure} of $\mathbb{P}$. A poset $\mathbb{P}$ is \emph{${<}\alpha$-strategically closed} if the second player has a winning strategy in the \emph{descending chain game} on $\mathbb{P}$ of length $\alpha$, which is a generalization of a usual Banach--Mazur game into transfinite length but the second player has the initiative in each limit inning (see \cite{Fo:gameba} or \cite{IY:games} for a precise definition). For an ordinal $\alpha$ which is either a limit or the successor of a limit, $\mathbb{P}$ is ${<}\alpha$-strategically closed if and only if \textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{CG}}^{<\alpha}(1)$ on $\mathbb{P}$ (it was proved in the case $\alpha=\omega+1$ by Jech and Veli\v{c}kovi\'{c} \cite{Je:multi,Vel:playful}, and in a general case by Ishiu in his unpublished paper~\cite{Ishiu:transgame}). It is unprovable in ZFC that if \textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{CG}}^{<\omega_1}(1)$ on $\mathbb{P}$ then $\mathbb{P}$ is ${<}\omega_1$-closed, for the following reason: It is known that $\mathbb{P}$ is ${<}\omega_1$-strategically closed if and only if $\mathbb{P}$ is ${<}(\omega+1)$-strategically closed (see \cite{Vel:square} or \cite{IY:games}), and it is known to be unprovable in ZFC that if $\mathbb{P}$ is ${<}(\omega+1)$-strategically closed then $\mathbb{P}$ is ${<}\omega_1$-closed (due to Jech and Shelah \cite{JeSh:cccba}). \end{rem} Here we list properties of a forcing notion $\mathbb{P}$ which are relevant to the results in this paper. $\operatorname{Fn}(\omega_1,2,\omega_1)$ denotes the poset which adjoins a Cohen subset of $\omega_1$ with countable conditions \cite{Ku:set}. The list is ordered \emph{stronger to weaker}. \begin{enumerate} \item $\mathbb{P}=\operatorname{Fn}(\omega_1,2,\omega_1)$. \item $\mathbb{P}$ is ${<}\omega_1$-closed. \item \textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{CG}}^{<\omega_1}(1)$ on $\mathbb{P}$. \item \textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{CG}}^{<\omega_1}({<}\aleph_0)$ on $\mathbb{P}$. \item \textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{CG}}^{\omega}({<}\aleph_0)$ on $\mathbb{P}$. \item \textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{CG}}^{\omega}(\aleph_0)$ on $\mathbb{P}$. \item $\mathbb{P}$ is proper. \end{enumerate} Now we turn to the games played on topological spaces. For a topological space $(X,\tau)$ and an ordinal $\alpha$, the \emph{game $\operatorname{\mathsf{G}}_1^{<\alpha}(\mathcal{O},\mathcal{O})$ on $(X,\tau)$} is played by two players \textsc{One}\xspace and \textsc{Two}\xspace for $\alpha$ innings as follows. In the inning $\beta<\alpha$, \textsc{One}\xspace chooses an open cover $\mathcal{U}_\beta$ of $X$ and then \textsc{Two}\xspace chooses $H_\beta\in\mathcal{U}_\beta$. \textsc{Two}\xspace wins in this game if there is $\gamma<\alpha$ such that, $\{H_\beta\mathchoice{:}{:}{\,:\,}{\,:\,} \beta<\gamma\}$ covers $X$. Note that \textsc{Two}\xspace does not win if just $\{H_\beta\mathchoice{:}{:}{\,:\,}{\,:\,} \beta<\alpha\}$ covers $X$. If $\alpha=\gamma+1$, We write $\operatorname{\mathsf{G}}_1^{\gamma}(\mathcal{O},\mathcal{O})$ instead of $\operatorname{\mathsf{G}}_1^{<\gamma+1}(\mathcal{O},\mathcal{O})$. We say a space $(X,\tau)$ \emph{has the Rothberger property} if, for every sequence $\langle\mathcal{U}_n\mathchoice{:}{:}{\,:\,}{\,:\,} n<\omega\rangle$ of open covers of $X$ there is an open cover $\{U_n\mathchoice{:}{:}{\,:\,}{\,:\,} n<\omega\}$ of $X$ such that $U_n\in\mathcal{U}_n$ for all $n<\omega$. It is easy to see that, if \textsc{One}\xspace does not have a winning strategy in the game $\operatorname{\mathsf{G}}_1^{\omega}(\mathcal{O},\mathcal{O})$ on $(X,\tau)$, then $(X,\tau)$ has the Rothberger property. The following theorem, which is due to Pawlikowski \cite{Pa:pointopen}, tells us that the converse also holds. \begin{thm}\label{thm:rothbergergame} A space $(X,\tau)$ has the Rothberger property if and only if \textsc{One}\xspace does not have a winning strategy in the game $\operatorname{\mathsf{G}}_1^{\omega}(\mathcal{O},\mathcal{O})$ on $(X,\tau)$. \end{thm} For a space $(X,\tau)$ and a forcing notion $\mathbb{P}$, we let $\tau^{\mathbb{P}}$ denote a $\mathbb{P}$-name representing the topology on $X$ generated by $\tau$ in a generic extension by $\mathbb{P}$. We say a forcing notion $\mathbb{P}$ \emph{destroys} a Lindel\"{o}f space $(X,\tau)$ if we have \[ \forcestext[\mathbb{P}]{ (\check{X},\tau^{\mathbb{P}})\text{ is not Lindel\"{o}f}}. \] A Lindel\"{o}f space $(X,\tau)$ is called an \emph{indestructibly Lindel\"{o}f} space if $(X,\tau)$ is not destroyed by any ${<}\omega_1$-closed poset. The equivalence $(1)\Leftrightarrow(2)$ in the following theorem is due to Scheepers and Tall \cite[Theorem~1]{ST:lindelof}. The equivalence $(2)\Leftrightarrow(3)$ is easily checked. \begin{thm}\label{thm:indlindelofgame} For a space $(X,\tau)$ the following are equivalent. \begin{enumerate} \item[\textup{(1)}] $(X,\tau)$ is an indestructibly Lindel\"{o}f space. \item[\textup{(2)}] $(X,\tau)$ is a Lindel\"{o}f space and \textsc{One}\xspace does not have a winning strategy in $\operatorname{\mathsf{G}}_1^{\omega_1}(\mathcal{O},\mathcal{O})$ on $(X,\tau)$. \item[\textup{(3)}] \textsc{One}\xspace does not have a winning strategy in $\operatorname{\mathsf{G}}_1^{<\omega_1}(\mathcal{O},\mathcal{O})$ on $(X,\tau)$. \end{enumerate} \end{thm} As a consequence of Theorem~\ref{thm:rothbergergame} and Theorem~\ref{thm:indlindelofgame}, we see the following fact \cite[Corollary~10]{ST:lindelof}. \begin{cor} A space with the Rothberger property is an indestructibly Lindel\"{o}f space. \end{cor} We say $X$ is a \emph{P-space} if every $\operatorname{\mathsf{G}}_{\delta}$-set in $X$ is an open set. It is known that a Lindel\"{o}f P-space has the Rothberger property (due to Galvin; see the following remark). \begin{rem} An open cover $\mathcal{U}$ of a space $X$ is an \emph{$\omega$-cover} if $X\notin\mathcal{U}$ and for every finite set $F\subseteq X$ there is a $U\in\mathcal{U}$ with $F\subseteq U$. An open cover $\mathcal{U}$ of $X$ is a \emph{$\gamma$-cover} if $\mathcal{U}$ is infinite and any infinite subset of $\mathcal{U}$ covers $X$. A space $X$ is called a \emph{$\gamma$-space} if, for every sequence $\langle\mathcal{U}_n\mathchoice{:}{:}{\,:\,}{\,:\,} n<\omega\rangle$ of $\omega$-covers of $X$ there is a $\gamma$-cover $\{U_n\mathchoice{:}{:}{\,:\,}{\,:\,} n<\omega\}$ of $X$ such that $U_n\in\mathcal{U}_n$ for all $n<\omega$. It is known that a $\gamma$-space has the Rothberger property. Galvin proved that a Lindel\"{o}f P-space is a $\gamma$-space (see \cite{GN:cx}, \cite[Theorem~47]{ST:lindelof}). \end{rem} Here we list properties of a topological space $X=(X,\tau)$ which are relevant to the results in this paper. The list is ordered \emph{weaker to stronger}. \begin{enumerate} \item $X$ is a Lindel\"{o}f space. \item $X$ is an indestructibly Lindel\"{o}f space (equivalently, \textsc{One}\xspace does not have a winning strategy in $\operatorname{\mathsf{G}}_1^{<\omega_1}(\mathcal{O},\mathcal{O})$ on $X$). \item $X$ has the Rothberger property (equivalently, \textsc{One}\xspace does not have a winning strategy in $\operatorname{\mathsf{G}}_1^{\omega}(\mathcal{O},\mathcal{O})$ on $X$). \item $X$ is a Lindel\"{o}f P-space. \end{enumerate} \begin{rem} Here we state facts about the topological property ``\textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{G}}_1^{<\omega_1}(\mathcal{O},\mathcal{O})$ on $X$'', which does not fit in the above list. Clearly, if \textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{G}}_1^{<\omega_1}(\mathcal{O},\mathcal{O})$ on $X$, then $X$ is an indestructibly Lindel\"{o}f space. Daniels and Gruenhage \cite{DG:pointopentype} showed that if $X$ is a hereditarily Lindel\"{o}f space then \textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{G}}_1^{<\omega_1}(\mathcal{O},\mathcal{O})$ on $X$. The real line $\mathbb{R}$ is a hereditarily Lindel\"{o}f space and so \textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{G}}_1^{<\omega_1}(\mathcal{O},\mathcal{O})$ on $\mathbb{R}$, whereas $\mathbb{R}$ does not have the Rothberger property. On the other hand, using results due to Scheepers and Tall \cite[Theorem 2 and Example 3]{ST:lindelof} we can see that, it is consistent with ZFC that there is a space with the Rothberger property on which \textsc{Two}\xspace does not have a winning strategy in $\operatorname{\mathsf{G}}_1^{<\omega_1}(\mathcal{O},\mathcal{O})$. \end{rem} \section{Proper forcing preserves P-spaces} We prove that a P-space remains a P-space after forcing with a proper forcing notion. \begin{prop}\label{prop:preservepspace} Suppose that a space $(X,\tau)$ is a P-space and $\mathbb{P}$ is a proper forcing notion. Then $\forcestext[\mathbb{P}]{ (\check{X},\tau^{\mathbb{P}})\text{ is a P-space}}$. \end{prop} \begin{proof} Fix a countable set $\{\dot{G}_n\mathchoice{:}{:}{\,:\,}{\,:\,} n<\omega\}$ of $\mathbb{P}$-names such that $ \forcestext[\mathbb{P}]{ \dot{G}_n\in\tau^{\mathbb{P}} } $ for all $n<\omega$. For each $n$, take a $\mathbb{P}$-name $\dot{\mathcal{T}}_n$ such that \[ \forcestext[\mathbb{P}]{ \dot{\mathcal{T}}_n\subseteq\check{\tau}\text{ and } \dot{G}_n=\bigcup\dot{\mathcal{T}}_n}. \] We are going to prove the following sentence. \[ \forcestext[\mathbb{P}]{ \forall x\in \check{X}\, \left[ x\in\bigcap_{n<\omega}\dot{G}_n \to \exists H\in\check{\tau}\,(x\in H \text{ and }H\subseteq\bigcap_{n<\omega}\dot{G}_n) \right] }, \] which implies that $\forcestext[\mathbb{P}]{ \bigcap_{n<\omega}\dot{G}_n\in\tau^{\mathbb{P}}} $. It suffices to show that, for $x\in X$ and $p\in \mathbb{P}$, if $p\forcestext[\mathbb{P}]{ \check{x}\in\bigcap_{n<\omega}\dot{G}_n}$, then there are $q\leq p$ and $H\in\tau$ such that $x\in H$ and $q\forcestext[\mathbb{P}]{ \check{H}\subseteq\bigcap_{n<\omega}\dot{G}_n}$. Fix $x\in X$, $p\in\mathbb{P}$ and assume $p\forcestext{\check{x}\in\bigcap_{n<\omega}\dot{G}_n}$. For each $n<\omega$, since we have $ p\forcestext[\mathbb{P}]{ \check{x}\in\dot{G}_n\text{ and } \dot{G}_n=\bigcup\dot{\mathcal{T}}_n} $, we can take a $\mathbb{P}$-name $\dot{T}_n$ such that $ p\forcestext[\mathbb{P}]{ \check{x}\in\dot{T}_n\text{ and } \dot{T}_n\in\dot{\mathcal{T}}_n} $. Note that we have $p\forcestext[\mathbb{P}]{ \dot{T}_n\in\check{\tau} \text{ and }\dot{T}_n\subseteq\dot{G}_n}$. By the properness of $\mathbb{P}$, we can choose $q\leq p$ and a countable set $\mathcal{C}\subseteq\tau$ so that $ q\forcestext[\mathbb{P}]{ \{\dot{T}_n\mathchoice{:}{:}{\,:\,}{\,:\,} n<\omega\}\subseteq\check{\mathcal{C}}} $. Note that $q\forcestext[\mathbb{P}]{ \forall n<\omega\,(\check{x}\in\dot{T}_n)}$. Let $H=\bigcap\{T\in\mathcal{C}\mathchoice{:}{:}{\,:\,}{\,:\,} x\in T\}$. Then $x\in H$ and, since $(X,\tau)$ is a P-space, $H\in\tau$ holds. Now we have \[ q\forcestext[\mathbb{P}]{ \check{H}=\bigcap{\{T\in\check{\mathcal{C}}\mathchoice{:}{:}{\,:\,}{\,:\,} \check{x}\in T\}} \subseteq\bigcap_{n<\omega}\dot{T}_n \subseteq\bigcap_{n<\omega}\dot{G}_n}, \] which concludes the proof. \end{proof} \section{The main result}\label{sec:main} In this section, we give a sufficient condition for a topological space $(X,\tau)$ and a forcing notion $\mathbb{P}$ to keep $(\check{X},\tau^{\mathbb{P}})$ having a certain game-theoretic property in the forcing extension. \begin{defin} For a topological space $X=(X,\tau)$, define a cardinal $p(X)$ by letting $ p(X) =\aleph_0+ \min\left( \{\size{\mathcal{G}}\mathchoice{:}{:}{\,:\,}{\,:\,} \mathcal{G}\subseteq\tau \text{ and } \bigcap\mathcal{G}\notin\tau \} \cup\{\size{\tau}^{+}\}\right) $. \end{defin} Note that $X$ is a P-space if and only if $p(X)\geq\aleph_1$. \begin{thm}\label{thm:main} Let $(X,\tau)$ be a topological space, $\mathbb{P}$ a forcing notion, $\alpha$ an ordinal and $\kappa=p(X)$. If \begin{enumerate} \item \textsc{One}\xspace does not have a winning strategy in $\operatorname{\mathsf{G}}_1^{<\alpha}(\mathcal{O},\mathcal{O})$ on $(X,\tau)$, and \item \textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{CG}}^{<\alpha}({<}\kappa)$ on $\mathbb{P}$, \end{enumerate} then \[ \forcestext[\mathbb{P}]{% \text{\textsc{One}\xspace does not have a winning strategy in } \operatorname{\mathsf{G}}_1^{<\alpha}(\mathcal{O},\mathcal{O}) \text{ on } (\check{X},\tau^{\mathbb{P}}) }. \] \end{thm} \begin{proof} Fix an enumeration of $\tau$, say $\tau=\{T_\xi\mathchoice{:}{:}{\,:\,}{\,:\,}\xi<\theta\}$ for some cardinal $\theta$. Suppose that $\dot{\sigma}$ is a $\mathbb{P}$-name such that \[ \forcestext[\mathbb{P}]{ \dot{\sigma} \text{ is a strategy for \textsc{One}\xspace in } \operatorname{\mathsf{G}}_1^{<\alpha}(\mathcal{O},\mathcal{O}) \text{ on } (\check{X},\tau^{\mathbb{P}}) }. \] Without loss of generality we may assume that, it is forced that the strategy $\dot{\sigma}$ suggests only open covers which consist of elements of $\tau$, since $\tau$ is a base of $\tau^{\mathbb{P}}$ in a generic extension, and taking refinements will not help \textsc{Two}\xspace win easier. Under this assumption, a sequence of initial moves for \textsc{Two}\xspace, played in a generic extension, against the strategy $\dot{\sigma}$ will be described in a form $\langle\check{T}_{\dot{\xi}_\beta}\mathchoice{:}{:}{\,:\,}{\,:\,}\beta<\delta\rangle$, where $\delta<\alpha$ and each $\dot{\xi}_\beta$ is a $\mathbb{P}$-name for an ordinal. We will prove the following statement: For any $p\in\mathbb{P}$, there are $q\leq p$, a sequence $\langle\dot{\xi}_\beta\mathchoice{:}{:}{\,:\,}{\,:\,}\beta<\alpha\rangle$ of $\mathbb{P}$-names for ordinals and $\gamma<\alpha$ such that \[ q\forcestext[\mathbb{P}]{ \forall\delta< \gamma \, \big(\check{T}_{\dot{\xi}_{\delta}}\in \dot{\sigma}(\langle\check{T}_{\dot{\xi}_\beta}\mathchoice{:}{:}{\,:\,}{\,:\,}\beta<\delta\rangle) \big) \text{ and } \bigcup\{\check{T}_{\dot{\xi}_{\delta}}\mathchoice{:}{:}{\,:\,}{\,:\,} \delta<\gamma\} =\check{X} }. \] This means that $\langle\check{T}_{\dot{\xi}_\beta}\mathchoice{:}{:}{\,:\,}{\,:\,}\beta<\alpha\rangle$ describes winning moves for \textsc{Two}\xspace against the given strategy $\dot{\sigma}$ for \textsc{One}\xspace in a generic extension. Fix $p\in\mathbb{P}$. By the assumption, \textsc{Two}\xspace has a winning strategy $\rho$ in the game $\operatorname{\mathsf{CG}}^{<\alpha}({<}\kappa)$ on $\mathbb{P}$ below $p$. We are going to define a strategy $\Sigma$ for \textsc{One}\xspace in the game $\operatorname{\mathsf{G}}_1^{<\alpha}(\mathcal{O},\mathcal{O})$ on $(X,\tau)$, which cannot be a winning strategy by the assumption. We construct $\Sigma$ by induction on $\delta<\alpha$. As an additional induction hypothesis we assume that, with each sequence $\langle H_\beta\mathchoice{:}{:}{\,:\,}{\,:\,}\beta<\delta\rangle$ describing \textsc{Two}\xspace's possible initial moves against $\Sigma$ before the inning $\delta$, a sequence $\langle\dot{\xi}_\beta\mathchoice{:}{:}{\,:\,}{\,:\,}\beta<\delta\rangle$ of $\mathbb{P}$-names for ordinals is associated. We will define \textsc{One}\xspace's move $\Sigma\left(\langle H_\beta\mathchoice{:}{:}{\,:\,}{\,:\,}\beta<\delta\rangle\right)$ in the inning $\delta$, and associate a $\mathbb{P}$-name $\dot{\xi}_\delta$ with \textsc{Two}\xspace's response $H_\delta$. Let $\dot{\mathcal{U}}^\delta$ be a $\mathbb{P}$-name such that $\forcestext[\mathbb{P}]{ \dot{\mathcal{U}}^\delta =\dot{\sigma}(\langle \check{T}_{\dot{\xi}_\beta} \mathchoice{:}{:}{\,:\,}{\,:\,}\beta<\delta\rangle)}$. Since we have $\forcestext[\mathbb{P}]{ \dot{\mathcal{U}}^\delta\subseteq\check{\tau} \text{ and } \bigcup\dot{\mathcal{U}}^\delta=\check{X} }$, for each $x\in X$ we can take a $\mathbb{P}$-name $\dot{\eta}_x^\delta$ for an ordinal so that \[ \forcestext[\mathbb{P}]{ \check{x}\in \check{T}_{\dot{\eta}_x^\delta} \text{ and } \check{T}_{\dot{\eta}_x^\delta}\in\dot{\mathcal{U}}^\delta }. \] For each $x\in X$, let $F_x =F\left(\langle\dot{\xi}_\beta\mathchoice{:}{:}{\,:\,}{\,:\,}\beta<\delta\rangle,x\right) =\rho\left(\langle\dot{\xi}_\beta\mathchoice{:}{:}{\,:\,}{\,:\,}\beta<\delta\rangle \conc\langle\dot{\eta}_x^\delta\rangle\right)$, and $G_x =G\left(\langle\dot{\xi}_\beta\mathchoice{:}{:}{\,:\,}{\,:\,}\beta<\delta\rangle,x\right) =\bigcap\{T_\xi\mathchoice{:}{:}{\,:\,}{\,:\,}\xi\in F_x\text{ and }x\in T_\xi\}$. Note that since $\size{F_x}<\kappa=p(X)$ and by the definition of $p(X)$, $G_x$ is an open set containing $x$. Now let \[ \Sigma\left(\langle H_\beta\mathchoice{:}{:}{\,:\,}{\,:\,}\beta<\delta\rangle\right) =\{G_x\mathchoice{:}{:}{\,:\,}{\,:\,} x\in X\}. \] Suppose that \textsc{Two}\xspace picks $H_\delta$ from the cover $\{G_x\mathchoice{:}{:}{\,:\,}{\,:\,} x\in X\}$ as a move in the inning $\delta$. We pick $x_\delta\in X$ such that $H_\delta=G_{x_\delta}$, and let $\dot{\xi}_\delta=\dot{\eta}_{x_\delta}^\delta$. This completes the induction step at $\delta$. Since $\Sigma$ is not a winning strategy, we can find a sequence $\langle H_\beta\mathchoice{:}{:}{\,:\,}{\,:\,}\beta<\alpha\rangle$ which describes \textsc{Two}\xspace's winning moves against $\Sigma$, and the associated sequence $\langle\dot{\xi}_\beta\mathchoice{:}{:}{\,:\,}{\,:\,}\beta<\alpha\rangle$ of $\mathbb{P}$-names of ordinals. For each $\delta<\alpha$, let $F_\delta=\rho(\langle\dot{\xi}_\beta\mathchoice{:}{:}{\,:\,}{\,:\,}\beta\leq\delta\rangle)$. Find $\gamma<\alpha$ such that $\{H_\beta\mathchoice{:}{:}{\,:\,}{\,:\,}\beta<\gamma\}$ covers $X$. Since $\rho$ is a winning strategy for \textsc{Two}\xspace in the game $\operatorname{\mathsf{CG}}^{<\alpha}({<}\kappa)$ on $\mathbb{P}$ below $p$, we can find $q\leq p$ such that $ q\forcestext[\mathbb{P}]{ \forall\delta<\gamma\, (\dot{\xi}_\delta\in \check{F}_\delta) }$. Fix $\delta<\gamma$. By the construction of the sequence $\langle\dot{\xi}_\beta\mathchoice{:}{:}{\,:\,}{\,:\,}\beta<\alpha\rangle$, we have $H_\delta =G\left(\langle\dot{\xi}_\beta\mathchoice{:}{:}{\,:\,}{\,:\,}\beta<\delta\rangle,x_\delta\right) $ and $\dot{\xi}_\delta=\dot{\eta}_{x_\delta}^\delta$ for a suitable $x_\delta\in X$. Note that $F_\delta =\rho(\langle\dot{\xi}_\beta\mathchoice{:}{:}{\,:\,}{\,:\,}\beta\leq\delta\rangle) =F\left(\langle\dot{\xi}_\beta\mathchoice{:}{:}{\,:\,}{\,:\,}\beta<\delta\rangle,x_\delta\right)$ and so $H_\delta =\bigcap\{T_\xi\mathchoice{:}{:}{\,:\,}{\,:\,}\xi\in F_\delta\text{ and }x_\delta\in T_\xi\} $. Also \[ \forcestext[\mathbb{P}]{ \check{x}_\delta\in\check{T}_{\dot{\xi}_\delta} \text{ and } \check{T}_{\dot{\xi}_\delta}\in \dot{\sigma} (\langle\check{T}_{\dot{\xi}_\beta}\mathchoice{:}{:}{\,:\,}{\,:\,}\beta<\delta\rangle) }. \] Since $q\forcestext[\mathbb{P}]{\dot{\xi}_\delta\in \check{F}_\delta}$ and by the definition of $H_\delta$, we have \[ q\forcestext[\mathbb{P}]{ \check{H}_\delta\subseteq\check{T}_{\dot{\xi}_\delta} }. \] Now we see $q\forcestext[\mathbb{P}]{ \forall\delta<\gamma\, (\check{H}_\delta\subseteq\check{T}_{\dot{\xi}_\delta}) }$, and since $\{H_\delta\mathchoice{:}{:}{\,:\,}{\,:\,}\delta<\gamma\}$ covers $X$, we have \[ q\forcestext[\mathbb{P}]{ \forall\delta<\gamma\, (\check{T}_{\dot{\xi}_{\delta}}\in \dot{\sigma} (\langle\check{T}_{\dot{\xi}_\beta}\mathchoice{:}{:}{\,:\,}{\,:\,}\beta<\delta\rangle) ) \text{ and } \bigcup\{\check{T}_{\dot{\xi}_{\delta}}\mathchoice{:}{:}{\,:\,}{\,:\,}\delta<\gamma\} =\check{X} }. \] This concludes the proof. \end{proof} \begin{rem} The reader might complain that, for $\delta\geq\gamma$, $q$ may not force $\check{T}_{\dot{\xi}_{\delta}}$ to be a possible move for \textsc{Two}\xspace. But it is unimportant, since the moves after the inning $\gamma$ do not affect the payoff and so \textsc{Two}\xspace may disregard $\check{T}_{\dot{\xi}_{\delta}}$'s and take any moves to follow the rule. \end{rem} A similar argument to the above proof yields the following corollary. An adaptation of the proof for the corollary is left to the reader. \begin{cor}\label{cor:twowins} Let $(X,\tau)$ be a topological space, $\mathbb{P}$ a forcing notion, $\alpha$ an ordinal and $\kappa=p(X)$. If \begin{enumerate} \item \textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{G}}_1^{<\alpha}(\mathcal{O},\mathcal{O})$ on $(X,\tau)$, and \item \textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{CG}}^{<\alpha}({<}\kappa)$ on $\mathbb{P}$, \end{enumerate} then \[ \forcestext[\mathbb{P}]{% \text{\textsc{Two}\xspace has a winning strategy in } \operatorname{\mathsf{G}}_1^{<\alpha}(\mathcal{O},\mathcal{O}) \text{ on } (\check{X},\tau^{\mathbb{P}}) }. \] \end{cor} \section{Consequences}\label{sec:conseq} Theorem~\ref{thm:main} together with Theorem~\ref{thm:indlindelofgame} yields the following consequence. \begin{cor}\label{cor:preserveindlindelof} Suppose that $(X,\tau)$ is an indestructibly Lindel\"{o}f space and $\mathbb{P}$ is a forcing notion such that \textsc{Two}\xspace has a winning strategy in the game $\operatorname{\mathsf{CG}}^{<\omega_1}({<}\aleph_0)$ on $\mathbb{P}$. Then \[ \forcestext[\mathbb{P}]{% (\check{X},\tau^{\mathbb{P}}) \text{ is an indestructibly Lindel\"{o}f space} }. \] \end{cor} Now consider the following three conditions on a Lindel\"{o}f space $(X,\tau)$. \begin{enumerate} \item[(1)] $(X,\tau)$ is not destroyed by $\operatorname{Fn}(\omega_1,2,\omega_1)$. \item[(2)] $(X,\tau)$ is indestructibly Lindel\"{o}f, that is, $(X,\tau)$ is not destroyed by any ${<}\omega_1$-closed forcing notion. \item[(3)] $(X,\tau)$ is not destroyed by any forcing notion $\mathbb{P}$ on which \textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{CG}}^{<\omega_1}({<}\aleph_0)$. \end{enumerate} Clearly $(3)\Rightarrow (2)\Rightarrow (1)$ holds. Tall pointed out (see \cite[Theorem~3]{Tall:lindelofpointsgd}) a result due to Shelah, which claims that $(1)\Rightarrow (2)$ holds. Corollary~\ref{cor:preserveindlindelof} tells us that $(2)\Rightarrow (3)$ holds, and hence these three conditions are all equivalent. In fact, we can also give a direct proof of $(1)\Rightarrow (3)$ by putting the argument of the proof of Theorem~\ref{thm:main} into Shelah's proof. Using Theorem~\ref{thm:main} and Theorem~\ref{thm:rothbergergame}, we see the following. \begin{cor}\label{cor:preserverothberger} Suppose that $(X,\tau)$ has the Rothberger property and $\mathbb{P}$ is a forcing notion such that \textsc{Two}\xspace has a winning strategy in the game $\operatorname{\mathsf{CG}}^{\omega}({<}\aleph_0)$ on $\mathbb{P}$. Then \[ \forcestext[\mathbb{P}]{% (\check{X},\tau^{\mathbb{P}}) \text{ has the Rothberger property} }. \] \end{cor} Let $\mathbb{B}(\kappa)$ denote the measure algebra on $2^{\kappa}$. Scheepers and Tall proved that, for any infinite cardinal $\kappa$, if $(X,\tau)$ has the Rothberger property, then $\forcestext[\mathbb{B}(\kappa)]{ (\check{X},\tau^{\mathbb{B}(\kappa)}) \text{ has the Rothberger property}}$ \cite[Theorem~15]{ST:lindelof}. It is known that \textsc{Two}\xspace has a winning strategy in the game $\operatorname{\mathsf{CG}}^{\omega}({<}\aleph_0)$ on $\mathbb{B}(\kappa)$ \cite{Je:moregame}, and hence Corollary~\ref{cor:preserverothberger} gives an alternate proof of their result. We also remark that Corollary~\ref{cor:preserverothberger} extends another result due to Scheepers and Tall, which claims that for a ${<}\omega_1$-closed forcing notion $\mathbb{P}$ if $(X,\tau)$ has the Rothberger property then $\forcestext[\mathbb{P}]{ (\check{X},\tau^{\mathbb{P}}) \text{ has the Rothberger property}}$ \cite[Theorem~21]{ST:lindelof}. As we mentioned in Section~\ref{sec:pre}, a Lindel\"{o}f P-space has the Rothberger property. Using this fact with Theorem~\ref{thm:countablechoiceproper}, Theorem~\ref{thm:rothbergergame}, Proposition~\ref{prop:preservepspace} and Theorem~\ref{thm:main}, we can deduce the following result. \begin{cor}\label{cor:preservelindelofpspace} Suppose that $(X,\tau)$ is a Lindel\"{o}f P-space and $\mathbb{P}$ is a forcing notion such that \textsc{Two}\xspace has a winning strategy in the game $\operatorname{\mathsf{CG}}^{\omega}(\aleph_0)$ on $\mathbb{P}$. Then \[ \forcestext[\mathbb{P}]{% (\check{X},\tau^{\mathbb{P}}) \text{ is a Lindel\"{o}f P-space} }. \] \end{cor} Now we can summarize these consequences of Theorem~\ref{thm:main} as in Table~\ref{table:summary}. The table is read as follows: ``A property of a topological space shown in a left-hand column is preserved under forcing extension by a forcing notion with the property shown in the corresponding right-hand column.'' \begin{table} \caption{Summary of consequences of the main result} \label{table:summary} \begin{center} \begin{tabular}{lll} \hline Topological spaces & Forcing notions & \\ \hline \hline (1)~Lindel\"{o}f & & \\ \hline & (1)~$\operatorname{Fn}(\omega_1,2,\omega_1)$ & \\ & (2)~${<}\omega_1$-closed & \\ & (3)~\textsc{Two}\xspace has a w.s. in $\operatorname{\mathsf{CG}}^{<\omega_1}(1)$ & \\ (2)~indestructibly Lindel\"{o}f & (4)~\textsc{Two}\xspace has a w.s. in $\operatorname{\mathsf{CG}}^{<\omega_1}({<}\aleph_0)$ & (\ref{cor:preserveindlindelof}) \\ \hline (3)~Rothberger & (5)~\textsc{Two}\xspace has a w.s. in $\operatorname{\mathsf{CG}}^{\omega}({<}\aleph_0)$ & (\ref{cor:preserverothberger}) \\ \hline (4)~Lindel\"{o}f P-space & (6)~\textsc{Two}\xspace has a w.s. in $\operatorname{\mathsf{CG}}^{\omega}(\aleph_0)$ & (\ref{cor:preservelindelofpspace}) \\ \hline & (7)~proper & \\ \hline \end{tabular} \end{center} \end{table} Before closing this section, we state a consequence of Corollary~\ref{cor:twowins}, which gives a sufficient condition for a forcing notion to preserve the topological property ``\textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{G}}_1^{<\omega_1}(\mathcal{O},\mathcal{O})$ on $X$''. \begin{cor}\label{cor:preservetwowins} Suppose that \textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{G}}_1^{<\omega_1}(\mathcal{O},\mathcal{O})$ on a topological space $(X,\tau)$ and $\mathbb{P}$ is a forcing notion such that \textsc{Two}\xspace has a winning strategy in the game $\operatorname{\mathsf{CG}}^{<\omega_1}({<}\aleph_0)$ on $\mathbb{P}$. Then \[ \forcestext[\mathbb{P}]{% \text{\textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{G}}_1^{<\omega_1}(\mathcal{O},\mathcal{O})$ on } (\check{X},\tau^{\mathbb{P}}) }. \] \end{cor} \section{Discussion} We will show that, under ZFC, the assumption ``\textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{CG}}^{\omega}({<}\aleph_0)$ on $\mathbb{P}$'' in Corollary~\ref{cor:preserverothberger} cannot be weakened to ``\textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{CG}}^{\omega}(\aleph_0)$ on $\mathbb{P}$''. We use the following famous result due to Laver~\cite{L:borel} (also found in \cite[Theorem~8.3.2]{BaJ:set}). Let $\mathbb{M}$ denote the Mathias forcing notion. \begin{thm}\label{thm:mathiassmz} Suppose that $X$ is an uncountable set of real numbers. Then \[ \forcestext[\mathbb{M}]{\check{X}\text{ does not have strong measure zero}}. \] \end{thm} It is easily checked that, if a forcing notion $\mathbb{P}$ satisfies Axiom A, then \textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{CG}}^{\omega}(\aleph_0)$ on $\mathbb{P}$. On the other hand, if \textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{CG}}^{\omega}({<}\aleph_0)$ on $\mathbb{P}$, then $\mathbb{P}$ is $\omega^\omega$-bounding by Theorem~\ref{thm:finitechoicebounding}. The Mathias forcing $\mathbb{M}$ satisfies Axiom A but is not $\omega^\omega$-bounding, and so \textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{CG}}^{\omega}(\aleph_0)$ on $\mathbb{M}$ but none in $\operatorname{\mathsf{CG}}^{\omega}({<}\aleph_0)$ on $\mathbb{M}$. Now assume CH and let $L$ be an uncountable Lusin set of real numbers. It is known that $L$ has the Rothberger property \cite{Ro:propertyc}. However, by Theorem~\ref{thm:mathiassmz} we have \[ \forcestext[\mathbb{M}]{ \check{L}\text{ does not have strong measure zero}}, \] and a set of real numbers with the Rothberger property has strong measure zero, which implies \[ \forcestext[\mathbb{M}]{ \check{L}\text{ does not have the Rothberger property}}. \] We do not know if the assumption ``\textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{CG}}^{<\omega_1}({<}\aleph_0)$ on $\mathbb{P}$'' in Corollary~\ref{cor:preserveindlindelof} can be weakened to ``\textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{CG}}^{\omega}({<}\aleph_0)$ on $\mathbb{P}$''. \begin{q} Can we find an indestructibly Lindel\"{o}f space $(X,\tau)$ and a forcing notion $\mathbb{P}$ which satisfy the following? \begin{enumerate} \item \textsc{Two}\xspace has a winning strategy in the game $\operatorname{\mathsf{CG}}^{\omega}({<}\aleph_0)$ on $\mathbb{P}$. \item $\forcestext[\mathbb{P}]{ (\check{X},\tau^{\mathbb{P}}) \text{ is not a Lindel\"{o}f space} }$. \end{enumerate} \end{q} \begin{rem} As we mentioned in Section~\ref{sec:conseq}, \textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{CG}}^{\omega}({<}\aleph_0)$ on the measure algebra $\mathbb{B}(\kappa)$ for any $\kappa$ (moreover, for any fixed $\alpha<\omega_1$, \textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{CG}}^{<\alpha}({<}\aleph_0)$ on $\mathbb{B}(\kappa)$). On the other hand, it is easy to find a winning strategy for \textsc{One}\xspace in $\operatorname{\mathsf{CG}}^{<\omega_1}({<}\aleph_0)$ on $\mathbb{B}(\kappa)$ (just note that any strictly decreasing sequence of real numbers has at most countable order type). Unfortunately, for any Lindel\"{o}f space $(X,\tau)$ we have $\forcestext[\mathbb{B}(\kappa)]{ (\check{X},\tau^{\mathbb{B}(\kappa)})\text{ is Lindel\"{o}f}}$ (see \cite{GJT:forcingnormality}). \end{rem} We do not know if the assumption ``\textsc{Two}\xspace has a winning strategy in $\operatorname{\mathsf{CG}}^{\omega}(\aleph_0)$ on $\mathbb{P}$'' in Corollary~\ref{cor:preservelindelofpspace} can be weakened to ``$\mathbb{P}$ is proper''. Let $\mathbb{CF}$ denote the poset which adjoins a closed unbounded subset of $\omega_1$ with finite conditions, which is due to Baumgartner \cite{Bau:applpfa}. It is known that $\mathbb{CF}$ is proper but \textsc{One}\xspace has a winning strategy in $\operatorname{\mathsf{CG}}^{\omega}(\aleph_0)$ on $\mathbb{CF}$. So it is natural to ask the following question. Note that, by Proposition~\ref{prop:preservepspace}, a P-space is still a P-space in a forcing extension by a proper poset. \begin{q}\label{q:destroylindelofp} Is there a Lindel\"{o}f P-space $(X,\tau)$ such that \[ \forcestext[\mathbb{CF}]{ (\check{X},\tau^{\mathbb{CF}})\text{ is not Lindel\"{o}f}}\;\text{?} \] \end{q} \subsection*{Acknowledgements} The author would like to thank Marion Scheepers for his helpful comments and discussion during this work.
\section{Introduction} When ultra-cold neutrons (UCNs) are trapped in the earth's gravitational field, their energy is quantized. In consequence, their probability density distribution exhibits their vertical modulation. The scale of this density modulation is calculated to be $(\hbar^2/2m_n^2g)^{1/3} \sim 6 \ \mathrm{\mu m}$. Observations of such quantum states have been presented in Refs. \cite{bib_Nesvizhevsky2002na,bib_Nesvizhevsky2003pr,bib_Nesvizhevsky2005ep}. We have proposed a more precise measurement using a pixel detector with an image magnification system \cite{bib_Sanuki}. Another pixel detectors for UCN have been developed recently, such as that reported in Ref. \cite{bib_timepix}. In this article we present the development of a pixel detector based on a commercial charge coupled device (CCD) covered with a neutron converter. By comparing $\mathrm{^{10}B}$ and $\mathrm{^6Li}$ as converter materials, we find that a $\mathrm{^6Li}$ converter produces energetic tritons which penetrate deep into the CCD in various directions, degrading the spatial resolution. Hence we conclude that $\mathrm{^{10}B}$ is an appropriate material for a neutron converter. \section{Detector design} The developed detector consists essentially of a CCD covered by a neutron converter. Charged particles produced via nuclear reaction in the converter are detected with the CCD. The choices of converter material and CCD are key for this detector. \subsection{Neutron converter} $\mathrm{^{10}B}$ and $\mathrm{^6Li}$ are chosen as test materials for the neutron converter, because of their large cross-sections with neutrons. The neutron absorption cross-sections for $\mathrm{^{10}B}$ and $\mathrm{^6Li}$ are $4.01 \times 10^3$ and $0.95 \times 10^3$ barn, respectively, for thermal neutrons ($v$ = 2,224 m/s). Neutrons react with $\mathrm{^{10}B}$ and $\mathrm{^6Li}$ in the following processes: \begin{align} \mathrm{n + ^{10}B }& \rightarrow \mathrm{\alpha\ (1.47\ MeV) + ^7Li\ (0.84\ MeV) + \gamma\ (0.48\ MeV)} &\mathrm{93.9\%} \label{eq_converterB}\\ & \rightarrow \mathrm{\alpha\ (1.78\ MeV) + ^7Li\ (1.01\ MeV)} & \mathrm\ {6.1\%} \notag \\ \mathrm{n + ^6Li} & \rightarrow \mathrm{\alpha\ (2.05\ MeV) + \mathrm{^3H}\ (2.73\ MeV)}& \mathrm{100\%} \label{eq_converterLi} \end{align} $\mathrm{^{10}B}$ has advantages over $\mathrm{^6Li}$ in that it has a larger cross section and a shorter range for converted charged particles in the CCD; $\mathrm{^{10}B}$ emits only short range particles. The ranges of $\alpha$(1.47 MeV), $\alpha$(1.78 MeV) and $\mathrm{^7Li}$(0.84 MeV, 1.01 MeV) in Si are 5, 6 and 2 $\mu$m, respectively. In contrast, $\mathrm{^6Li}$ emits $\alpha$ particles and tritons with ranges of 7 and 40 $\mu$m. Long range tritons degrade the spatial resolution, as discussed in Ref. \cite{bib_Sanuki}. \subsection{CCD sensor} A CCD is an ideal device for UCN detection because of its high spatial resolution and its ability to capture data in real time. Since a CCD cannot directly detect neutral particles, a neutron converter to create charged particles must be attached to the front of the device. A back-thinned type CCD is chosen to avoid a large insensitive volume in front of the active volume of the CCD, which would prevent converted charged particles entering the active volume or degrade the spatial resolution. We use a commercial CCD detector, HAMAMATSU S7170-0909, in the low dark current read out mode. The specifications are shown in Table \ref{CCD:spec}. Generally, CCD might have white-spot pixels induced by radiation damage. Only 2 pixels have become noisy after 60 million neutron irradiation per 512$\times$512 pixels during the neutron beam tests mentioned below. \begin{table}[htdp] \caption{Specifications of the HAMAMATSU S7170-0909 ($\mathrm{T_a = 25}$\kern-.2em\r{}\kern-.3em C)} \begin{center} \begin{tabular}{ll} \hline Parameter & \\ \hline Active Area & 12.288 $\times$ 12.288 mm \\ Number of Pixels & 512 $\times$ 512 \\ Pixel Size & 24 $\times$ 24 $\mu$m \\ Frame Rate & 0.9 frames/s \\ Full Well Capacity (Vertical) & 300 k$e^-$ \\ Full Well Capacity (Horizontal) & 600 k$e^-$ \\ Dark Current Max. $\mathrm{0^\circ C}$ & 600 $e^-$/pixel/s \\ Readout Noise & 8 $e^-$rms \\ \hline \end{tabular} \end{center} \label{CCD:spec} \end{table} \subsection{Fabrication of a converter layer} The neutron converter is attached directly to the surface of a CCD to minimize the distance between the positions of an incident neutron and a detected charged particle \cite{bib_Sanuki}. The converter layer is mounted on the CCD surface by vacuum evaporation at Kyoto University Research Reactor Institute. The pressure of the evaporation chamber is around $\mathrm{10^{-3} \ Pa}$ and the evaporation rate is about 1 \AA $\mathrm{/s}$ for each material. The facility used is described in Ref. \cite{bib_KURRI}. Converter layers of 46 $\mathrm{\mu g \,cm^{-2}}$ ($\mathrm{220\ nm}$) and 11 $\mathrm{\mu g \, cm^{-2}}$ ($\mathrm{230\ nm}$) for $\mathrm{^{10}B}$ and $\mathrm{^6Li}$, respectively, are attached to the CCD surface. To form a firm layer, a 9 $\mathrm{\mu g \, cm^{-2}}$(20\ nm) Ti layer is directly deposited on the CCD surface and the converter layer is mounted on the Ti layer. To repel moisture from the air, the outer surface of the converter layer is covered with a 9 $\mathrm{\mu g \, cm^{-2}}$(20\ nm) Ti layer. (Note that we incorrectly reported the amount of Ti and Li layer in our previous report \cite{bib_Sanuki}.) Ti has good characteristics for this use, because of its chemical stability, adhesion and negative potential for neutrons. The stability of these Ti--$\mathrm{^{10}B}$--Ti and Ti--$\mathrm{^6Li}$--Ti structures have allowed our developed detector to work reliably for more than three years. \section{Performance of a CCD sensor} \label{pef.:CCD} We investigate the CCD response to charged particles using $\alpha$ particles from $\mathrm{^{241}Am}$. The CCD and the $\mathrm{\alpha}$ source are put in a chamber filled with dry $\mathrm{N_2}$ gas, with the distance between them fixed at $\mathrm{123\ mm}$. The $\mathrm{N_2}$ gas pressure is varied to change the energy of the $\mathrm{\alpha}$ particles at the surface of the CCD. The energy to create an electron--hole pair in Si is $\mathrm{3.65\ eV}$, and hence an energy deposit of $\mathrm{1\ MeV}$, for instance, creates about 300,000 electron--hole pairs. Electrons are collected at the anode and stored in a charge capacitance. During storage, electrons spread to the adjacent pixels and make a cluster. Fig. \ref{signals} shows typical cluster shapes made by $\mathrm{\alpha}$ particles of energy 1 MeV and 4 MeV. For $\alpha$ particles with energy lower than 2 MeV, the created electrons diffuse isotropically in the vertical and horizontal directions. For more energetic particles, higher than 2 MeV, the cluster shapes become anisotropic. This is because so many electron--hole pairs are created that they overflow the full well capacity. As shown in Table \ref{CCD:spec}, the vertical full well capacity is lower than the horizontal capacity. Overflowed electrons spread only in the vertical direction and make an anisotropic cluster. \begin{figure}[htdp] \begin{center} \resizebox{80mm}{!}{\includegraphics{fig1}} \caption{Typical signals made by $\alpha$ particles of energy (a) 1 MeV and (b) 4 MeV. An $\alpha$ particle of energy lower than 2 MeV creates an isotropic cluster, while an $\alpha$ particle of energy higher than 2 MeV creates an anisotropic cluster spread in the vertical direction.} \label{signals} \end{center} \end{figure} Considering the anisotropy of the full well capacity, the signal region is determined to be $\mathrm{7 \ pixels}$ (horizontal) $\times $ $\mathrm{11 \ pixels}$ (vertical) around the peak pixel, collecting all the created electrons. More than 99\% of the diffused electrons are stored in this region. Fig. \ref{calibration} shows the energy calibration curve of the total charge in the signal region to the energy of $\alpha$ particles. Linearity is confirmed below 4 MeV. $\alpha$ particles of energy more than 4 MeV are so energetic that they can penetrate a CCD detection volume thickness of 20 $\mathrm{\mu m}$. \begin{figure}[tdp] \begin{center} \resizebox{100mm}{!}{\includegraphics{fig2.eps}} \caption{Energy calibration curve. The total charge in the signal region is proportional to the energy of the $\alpha$ particle below 4 MeV.} \label{calibration} \end{center} \end{figure} When the neutron rate is 100 events/frame, more than one event is observed inside one signal region. \section{Neutron beam test} The performance of neutron detectors with a $\mathrm{^{10}B}$ converter and a $\mathrm{^6Li}$ converter are tested with neutrons from reactors. \subsection{Neutron detection efficiency} The neutron detection efficiency is measured using UCNs supplied at the PF2 beam line of Institut Laue--Langevin in France. Fig. \ref{energy} shows the energy spectra measured by CCDs with $\mathrm{^{10}B}$ and $\mathrm{^6 Li}$ converters. Peaks corresponding to the converted charged particles (see Eq. \ref{eq_converterB}, \ref{eq_converterLi}) are seen in the spectra. The peak energy is slightly lower than the converted charged particle energy because of the energy loss in the converter itself. A peak around $\mathrm{0.9\ MeV}$ in the energy spectrum for the $^6 \mathrm{Li}$ converter is caused by tritons which are so energetic that they penetrate through the sensitive volume of the CCD \cite{bib_Sanuki}. Tritons emitted in a large incident angle at the surface of the CCD deposit all their energy and stop in the CCD. \begin{figure}[htdp] \begin{center} \resizebox{100mm}{!}{\includegraphics{fig3.eps}} \caption{Energy spectra observed with CCDs with (a) $\mathrm{^{10}B}$ converter and (b) $\mathrm{^6Li}$ converter} \label{energy} \end{center} \end{figure} To select neutron events, the minimum energy cut is set to 0.3 MeV for the $^{10} \mathrm{B}$ converter and to 0.8 MeV for the $^6 \mathrm{Li}$ converter. The incident neutron flux is measured with a $^3\mathrm{He}$ gas counter with an efficiency greater than 83\%. The detection efficiency is measured to be 44.1\,$\pm$\,0.7(stat.)\,$\pm$\,0.8(syst.)\% for the $^{10} \mathrm{B}$ converter and 21.0\,$\pm$\,0.8(stat.)\,$\pm$\,0.5(syst.)\% for the $^6 \mathrm{Li}$ converter, normalized by the number of neutrons measured by the $\mathrm{^3He}$ gas counter. The main source of the systematic error is uncertainty in the initial flux of neutrons. The neutron conversion efficiency for the $\mathrm{^{10}B}$ converter is higher than that for the $\mathrm{^6Li}$ as expected from their cross-sections. \subsection{Spatial resolution} The electrons spread over nearby pixels and make a cluster, as shown in Fig. \ref{signals}. The barycenter of charge is a good estimation of the neutron incident position. We measure the spatial resolution of the CCD using reference patterns made with a neutron absorber. The reference pattern is placed just in front of the CCD and is irradiated by neutrons. The distance between the reference pattern and the CCD surface is as close as $\mathrm{150\ \mu m}$. The reference pattern and the CCD are exposed to cold neutrons (CNs) at the MINE2 beam line of the research reactor JRR-3. The monochromatic neutron beam has a wavelength of 0.88 nm and a bandwidth of 2.7\% in full width at half maximum. CNs are suitable for evaluating the spatial resolution because their beam divergence is as small as $1\times10^{-3}$ radian. \subsubsection{Comparison between $\mathrm{^{10}B}$ and $\mathrm{^6Li}$ converters} We have reported that a CCD sensor covered with a $\mathrm{^6Li}$ converter can determine a neutron incident position with a precision of $\mathrm{\mathrm{5.3\ \mu m}}$ \cite{bib_Sanuki}. The spatial resolution of a CCD with a $\mathrm{^{10}B}$ converter is compared to that of a $\mathrm{^6Li}$ converter using the same reference pattern as used in our previous report. The reference pattern is made of a gadolinium foil with a thickness of $\mathrm{25\ \mu m}$. This thickness is sufficient to fully absorb CNs. The 144 fine spoke slits shown in Fig. \ref{laser_pattern}(a) are formed by an excimer laser by Laserx Co. Ltd. Each spoke slit has four trapezoidal holes along the radius, as shown in Fig. \ref{laser_pattern}(b) \cite{bib_Sanuki}. \begin{figure}[htdp] \begin{center} \resizebox{100mm}{!}{\includegraphics{fig4.eps}} \caption{(a) Reference pattern created by a laser process. (b) Enlargement of one spoke. The black areas are hole positions.} \label{laser_pattern} \end{center} \end{figure} The spatial resolution depends on the range and deposit energy of the charged particle created by nuclear reaction in the converter. To investigate differences in the spatial resolution of charged particles, the energy range is selected to identify different charged particles (see Table \ref{different_resolution}). \begin{figure}[htdp] \begin{center} \resizebox{80mm}{!}{\includegraphics{fig5.eps}} \caption{Hit position distribution measured with a pattern created by a laser process} \label{scat} \end{center} \end{figure} Fig. \ref{scat} shows the hit position distributions calculated from the barycenter of each cluster. All the events are plotted in an area one-twelfth of the total area, assuming rotational symmetry of the event distribution. For the stopped tritons from the $^6 \mathrm{Li}$ converter, the innermost slit cannot be distinguished. This is because the stopped tritons span a long range inside the CCD, as mentioned above. On the other hand, a clear pattern can be seen for other charged particles. These figure indicate that the spatial resolution is much better than the smallest pattern distance of $\mathrm{30 \ \mu m}$. In order to estimate the spatial resolution more precisely, the hit position distribution at the edge of the slit is investigated. The spatial resolution is evaluated by fitting an error function with a constant background expressed as \begin{equation} f(x) = p_1 \cdot \mathrm{erf} \left(\frac{x}{\sqrt{2} \sigma} +p_2 \right) + p_3, \end{equation} where $x$ is the distance from the edge, $\sigma$ corresponds to the spatial resolution and $p_1$, $p_2$ and $p_3$ are free fitting parameters. The fitting result for $\alpha$(1.47 MeV) from the $\mathrm{^{10}B}$ converter is shown in Fig. \ref{fit} as an example. The obtained resolutions ($\sigma$ of the error function) are shown in Table \ref{different_resolution}. The spatial resolution of the stopped tritons is too poor to be evaluated by the fitting. \begin{figure}[htdp] \begin{center} \resizebox{100mm}{!}{\includegraphics{fig6.eps}} \caption{Distance between the hit position and the edge of slit. Only $\alpha$(1.47 MeV) events from the $\mathrm{^{10}B}$ converter are displayed.} \label{fit} \end{center} \end{figure} We find that the $\mathrm{^{10}B}$ converter has higher efficiency and better spatial resolution than the $\mathrm{^6Li}$ converter. We conclude that $\mathrm{^{10}B}$ is a more suitable converter material for a UCN detector. \begin{table} \caption{resolution of each charged particle} \label{different_resolution} \begin{center} \begin{tabular}{cccc} \hline converter &energy range (MeV)&particle& $\sigma$ ($\mathrm{\mu m}$) \\ \hline & 0.2 -- 0.7 & $\mathrm{^7 Li}$ & 5.7 $\pm$ 0.2\\ $\mathrm{^{10}B}$ & 0.9 -- 1.35 & $\mathrm{ \alpha}$ & 5.5 $\pm$ 0.2 \\ & 1.35 -- 1.7 & $\mathrm{ \alpha}$ & 7.6 $\pm$ 1.0 \\ \hline & 0.8 -- 1.2 &$\mathrm{^3He (penetrate)}$& 6.7 $\pm$ 0.6 \\ $\mathrm{^6Li}$ & 1.4 -- 2.0 & $\mathrm{\alpha}$ & 5.3 $\pm$ 0.3 \\ & 2.2 -- 2.7 & $\mathrm{^3He (stop)} $ & --- \\ \hline \end{tabular} \end{center} \end{table} \subsubsection{Precise evaluation of spatial resolution} The reference pattern used in the above test does not have perfectly straight edges, probably because the gadolinium foil is too thick to form sharp edges. The variations from a straight slit edge are observed to be 1--2 $\mathrm{\mu m}$ by optical microscopy. The spatial resolution evaluated using the reference pattern (Table \ref{different_resolution}) includes this effect \cite{bib_Sanuki}. To evaluate the intrinsic spatial resolution in our detector, we fabricated a more precise reference pattern at the Center for Integrated Nano Technology Support of Tohoku University. Fig. \ref{Gd_pattern} shows the layout of the reference Gd pattern. The pattern consists of stripes that have different widths of 1--40 $\mathrm{\mu m}$. The pattern is fabricated by a damascene process, usually used for semiconductor devices \cite{bib_damascene}. The damascene process is shown in Fig. \ref{damascene} \begin{figure}[thbp] \begin{center} \resizebox{100mm}{!}{\includegraphics{fig7.eps}} \caption{Reference pattern created by damascene process. The black areas are Gd-less areas.} \label{Gd_pattern} \end{center} \end{figure} \begin{figure}[bhtp] \begin{center} \resizebox{100mm}{!}{\includegraphics{fig8.eps}} \caption{Damascene process: 1 depositing a hard mask layer on a Si substance by photo lithography, 2 making trenches on the Si with plasma reactive ion etching, 3 removing the mask layer, 4 sputtering gadolinium onto the trenches, 5 removing excess gadolinium by chemical-mechanical polishing } \label{damascene} \end{center} \end{figure} This process produces an edge to the pattern that is sufficiently sharp. The thickness of the gadolinium is 2--5 $\mathrm{\mu m}$, which corresponds to an absorption probability of 75--97\% for CNs. Fig. \ref{5um_trench} shows the 5 $\mu$m-wide trench after polishing. \begin{figure}[htdp] \begin{center} \resizebox{100mm}{!}{\includegraphics{fig9.eps}} \caption{A 5 $\mu$m-wide trench made by damascene process} \label{5um_trench} \end{center} \end{figure} Using this striped pattern, the differences in the spatial resolution with respect to the read out directions are measured. Fig. \ref{scatterHV} shows the hit position distribution measured with the pattern, and Fig. \ref{resolution} shows the their projections onto the horizontal and vertical direction. The spatial resolution is evaluated by fitting. The fitting function is expressed by the sum of an error function and a complementary error function: \begin{equation} f(x) = p_1\cdot \sum_{n=1}^{12} \left[ \mathrm{erfc} \left(\frac{x-d_n}{\sqrt{2} \sigma}+p_2 \right) + \mathrm{erf} \left(\frac{x-(d_n+w_n)}{\sqrt{2} \sigma}+p_2 \right) \right] + p_3, \end{equation} where $x$ is the hit position, $\sigma$ is the spatial resolution, $d_n$ is the distance from the initial position of the pattern to the $n_{th}$ stripe, $w_n$ is the width of the $n_{th}$ stripe and $p_1$, $p_2$ and $p_3$ are free fitting parameters. The 12 widest stripes, whose widths are 5--40 $\mathrm{\mu m}$, are used for this fitting. This is because gadolinium cannot enter trenches narrower than $\mathrm{5\ \mu m}$ in the spattering process. \begin{figure}[htdp] \begin{center} \resizebox{80mm}{!}{\includegraphics{fig10.eps}} \caption{Hit position distribution measured with a pattern created by a damascene process. The striped pattern is placed along the vertical direction (a) and the horizontal direction (b) } \label{scatterHV} \end{center} \end{figure} \begin{figure}[htdp] \begin{center} \resizebox{100mm}{!}{\includegraphics{fig11.eps}} \caption{Projected hit position: (a) horizontal projection of Fig.\ref{scatterHV}(a) and (b) vertical projection of Fig.\ref{scatterHV}(b). The spatial resolutions are evaluated to be 2.9\,$\pm$\,0.1 $\mathrm{\mu m}$ (horizontal) and 4.3\,$\pm$\,0.2 $\mathrm{\mu m}$ (vertical) in terms of sigma of error function} \label{resolution} \end{center} \end{figure} In the fitting result, the horizontal and vertical resolutions are evaluated to be 2.9\,$\pm$\,0.1 $\mathrm{\mu m}$ and 4.3\,$\pm$\,0.2 $\mathrm{\mu m}$, respectively. These values correspond to 6.9\,$\pm$\,0.2 $\mathrm{\mu m}$ and 9.9\,$\pm$\,0.3 $\mathrm{\mu m}$ of the full width at half maximum of the line spread functions. This difference is caused by the difference in the full well capacity. The measured spatial resolution is much finer than the pixel size of 24 $\mathrm{\mu m}$. \section{Summary} We developed a pixel detector for UCN which has high spatial resolution with temporal information. The detector is based on a commercial CCD on which a suitable neutron converter is directly evaporated. $\mathrm{^{10}B}$ and $^6\mathrm{Li}$ are tested as neutron converter materials. We conclude that $\mathrm{^{10}B}$ has better properties as a neutron converter material. Using $\mathrm{^{10}B}$, the UCN detection efficiency is 44.1\,$\pm$\,1.1\%. Our UCN detector can measure the UCN incident position with a precision of 2.9\,$\pm$\,0.1 $\mathrm{\mu m}$ (horizontal) and 4.3\,$\pm$\,0.2 $\mathrm{\mu m}$ (vertical) in terms of sigma of the error functions. \section{Acknowledgment} We are indebted to Prof. H. M. Shimizu of the High Energy Accelerator Research Organization (KEK) for constructive comments and encouragement. We would also like to thank Prof. P. Geltenbort of Institut Laue--Langevin (ILL) and Prof. S. Tanaka of Tohoku University for their technical support and encouragement. A part of this study was supported by the Center for Integrated Nanotechnology Support at Tohoku University and also by Nanotechnology Network Project of the Ministry of Education, Culture, Sports, Science, and Technology (MEXT) of the Japanese Government. This study was also supported by a Grant-in-Aid for Scientific Research (KAKENHI: 18409056, 20340050) from the Japan Society for the Promotion of Science and Research Fellowships of the Japan Society for the Promotion of Science for Young Scientists (19.4404).
\section{Introduction} A packing is defined as a set of nonoverlapping objects arranged in a space of given dimension. Packings of identical nonoverlapping spheres in $d$-dimensional Euclidean space ${\mathbb R}^d$ have been employed in condensed matter and materials physics as models for the structures of a diverse range of substances from crystals and colloids to liquids, amorphous solids and glasses \cite{CLPCMP1995,HMTSL2006,ZallenPAS1983}. In structural biology, molecular dynamics simulations of interactions between large numbers of molecules employ chains of nonoverlapping spheres as models for various biological structures such as proteins and lipids \cite{DND2007a,Dokholyan2006a,DD2005a}. In part due to the ability of these conceptually simple models to describe many of the fundamental characteristics of more complex substances, understanding the properties of sphere packings has also long been an area of interest in mathematics (for example, see \cite{CS1995a}). However, solving even some of the most basic of mathematical problems has proved challenging. For example, a proof of the Kepler conjecture, a proposition stating that the face-centered cubic lattice is the densest possible arrangement of spheres for $d=3$, has only recently emerged \cite{Hales2005a}. Furthermore, the kissing number $K_d$, or number of identical $d$-dimensional nonoverlapping spheres that can simultaneously be in contact with (kiss) a central sphere, was until recently only known rigorously for $d=1\!-\!3$, $8$ and $24$ \cite{CSSPLG1999}, though Musin \cite{Musin2008a} has now proved the $d=4$ case ($K_4 = 24$). One sphere packing problem that has not been generally addressed for an arbitrary number of spheres is that of finding the maximally dense (optimal) packing(s) of $N$ identical $d$-dimensional nonoverlapping spheres near (local to) an additional fixed central sphere such that the greatest radius $R$ from any of the surrounding spheres' centers to the center of the fixed sphere is minimized. This problem is called the densest local packing (DLP) problem \cite{HST2009b}. There is a single minimized greatest radius, denoted by $R_{min}(N)$, for each $N$ in the DLP problem in ${\mathbb R}^d$, though generally for each $N$ there may be multiple distinct packings that achieve this radius. Figure \ref{DLP15pic} depicts a conjectured optimal packing, belonging to point group $D_{5h}$ \cite{CottonCAGT1990}, for the DLP problem for $N=15$, $d=2$, with $R_{min}(15) = 1.873123\dots$ \cite{endnote3}. \begin{figure}[ht] \centering \includegraphics[width = 3.0in,viewport = 5 5 460 465,clip]{fig1} \caption{A conjectured DLP optimal packing (point group $D_{5h}$) for $N=15$, $d=2$, $R_{min}(15) = 1.873123\dots$, with encompassing sphere of radius $R_{min}(15) + 0.5 = 2.373123\dots$.} \label{DLP15pic} \end{figure} In various limits, the densest local packing problem encompasses both the kissing number and (infinite) sphere packing problems. The former is a special case of the DLP problem in that $K_d$ is equal to the greatest $N$ for which $R_{min}(N) = 1$, and the latter is equivalent to the DLP problem in the limit that $N \rightarrow \infty$. The equivalence of the latter problem may be explained by observing that in the limit as $N \rightarrow \infty$, the boundary of radius $R_{min}(N) \rightarrow \infty$, and that in this limit the ratio of the number of spheres within a fixed finite distance of the boundary to the number in the bulk is zero. The densest local packing problem is relevant to the realizability of functions that are candidates to be the pair correlation function of a packing of identical spheres. For a statistically homogeneous and isotropic packing, the pair correlation function is denoted $g_2(r)$; it is proportional to the probability density of finding a separation $r$ between any two sphere centers and normalized such that it takes the value of unity when no spatial correlations between centers are present. Specifically, no function can be the pair correlation function of a point process (where a packing of spheres of unit diameter is a point processes in which the minimum pair separation distance is unity) unless it meets certain necessary, but generally not sufficient, conditions known as realizability conditions \cite{Lenard1975a,TS2002a,KLS2007a}. Two of these conditions that appear to be particularly strong for the realizability of sphere packings \cite{TS2006a} are the nonnegativity of $g_2(r)$ and its corresponding structure factor $S(k)$, where \begin{equation} S(k) = 1 + \rho \tilde{h}(k) \label{structFactCond} \end{equation} with number density $\rho$ and \begin{equation} \tilde{h}(k) = (2\pi)^{d/2}\int_0^{\infty}r^{d-1}h(r)\frac{J_{d/2-1}(kr)}{(kr)^{d/2-1}}dr \label{dDimFourier} \end{equation} the $d$-dimensional Fourier transform of the total correlation function $h(k) \equiv g_2(r)-1$, with $J_{\nu}(x)$ the Bessel function of the first kind of order $\nu$. The $g_2$-invariant process of Torquato and Stillinger \cite{TS2002a} is a method to maximize the number density $\rho$ associated with the structure factor $S(k)$ of a given parameterized family of test functions, where each function in the family is a candidate to be the pair correlation function of a statistically homogeneous and isotropic packing of spheres. In the $g_2$-invariant process, the problem of finding the maximal achievable $\rho$ is posed as an optimization problem: maximize $\rho$ over the parameters subject to the nonnegativity of the test function and its corresponding structure factor. This process could be improved by the addition of further realizability conditions on the pair correlation function, assuming that these further conditions included information beyond that incorporated in the two nonnegativity conditions discussed above. Knowledge of the maximal number of sphere centers that may fit within radius $R$ from an additional fixed sphere center, where that maximal number is equal to the greatest $N$ in the DLP problem for which $R_{min}(N) \leq R$, may be employed to construct an additional realizability condition on $g_2(r)$. As was discussed in previous papers \cite{HST2009a,HST2009b}, this realizability condition has been shown to encode information not included in the nonnegativity conditions on pair correlation functions and their corresponding structure factors alone. The DLP problem may be alternatively stated as the problem of finding the densest packing of $N$ identical nonoverlapping spheres of unit diameter near an additional fixed sphere, where number density $\rho$ is measured over the volume enclosed by an \textit{encompassing sphere} of radius $R+0.5$ (see Fig. \ref{DLP15pic}) centered on the fixed sphere. We note that for $N$ spheres of unit diameter, the number density $\rho$ of the packing is linearly proportional (by a constant that varies only with dimension) to the packing fraction $\phi(R+0.5)$, the fraction of the volume of the encompassing sphere covered by the spheres of unit diameter. As will be discussed in detail later, the maximal infinite-volume packing fraction $\phi_*^{\infty}$ of identical nonoverlapping spheres in $d$ dimensions may be bounded from above by employing a specific definition of local packing fraction for a given number $N$ of spheres. For small numbers of spheres ($N \leq 1000$) in low dimensions ($d \leq 10$), a algorithm combining a nonlinear programming method with a stochastic search of configuration space can be employed on a personal computer to find solutions to the DLP problem. Using such an algorithm, the details of which are outlined in Appendix \ref{algorithm}, we find and present putative DLP optimal packings and their corresponding $R_{min}(N)$ in ${\mathbb R}^2$ for $N = 1$ to $N=109$, and for the values of $N$ corresponding to full shells of the triangular lattice from $N=120$ to $N=348$. Though we recognize that the putative optimal packings found by our algorithm are not rigorously proved to be optimal, we analyze each configuration of $N$ spheres under the assumption that it is a global minimum of the DLP problem. This assumption of optimality is supported by the proved robustness of the algorithm in recovering the known and strongly conjectured global minima of the DLP problem (e.g., the kissing numbers for $d=1\!-\!4$ and for $d=8$, the curved hexagonal packings for $d=2$; $N=18$, $36$, $60$, $90$ and $126$ \cite{LG1997a}) and by repeated testing. The aforementioned realizability condition on the pair correlation function is valid whether or not the putative optimal packings we have found are indeed global minima, as global minima simply provide the most restrictive realizability condition. However, the upper bound on the maximal density of an infinite sphere packing requires knowledge of proved optimal $R_{min}(N)$ to be rigorously correct, though we have found in practice that for $d=2$ and over the range of $N$ tested that our putative bound is valid. With regard to this finding and the proved robustness of the algorithm over the range of $N$ studied, in the following sections we refer to all DLP packings and $R_{min}(N)$ presented as optimal. In Sec. II, we discuss the realizability condition that results from knowledge of a finite number of $R_{min}(N)$ in a space ${\mathbb R}^d$ of arbitrary dimension $d$. We present the condition derived from knowledge of $R_{min}(N)$ for $N=1$ to $N=109$ in ${\mathbb R}^2$ and compare the $R_{min}(N)$ values to the shell distances in a triangular lattice of $N$ disks. In Sec. III, we construct a logical argument to prove the validity of the aforementioned upper bound, and we present the $d=2$ upper bounds derived from our method for selected $N$ from $N=6$ to $N=348$. In Sec IV, we present $d=2$ optimal packings and their corresponding $R_{min}(N)$ for selected values of $N$ from $N=10$ to $N=348$. We analyze the optimal packings presented and discuss their symmetry characteristics, noting that there is significant variability in both the configurations and symmetry elements of optimal packings over the range of $N$ studied. In Sec. V, we summarize our results and findings, and we discuss some of the implications of our work. In a sequel to this paper, we will present and analyze DLP optimal packings and their corresponding $R_{min}(N)$ for the $d=3$ case over a similar range of $N$. We will compare $R_{min}(N)$ values to shell distances in Barlow packings \cite{Barlow1883a}, where the Barlow packings, of which the best-known are face centered cubic (FCC) and hexagonal close packed (HCP) arrangements, all individually achieve the maximal infinite-volume packing fraction for $d=3$, $\phi_*^{\infty} = \pi/\sqrt{18} = 0.740481\dots$. The coordinates for and images of DLP optimal packings and values for $R_{min}(N)$ over the entire range of $N$ studied can be found on the authors' website \cite{website}. \section{Pair correlation function realizability and $Z_{max}(R)$} The realizability condition on $g_2(r)$ results from a relation between an upper bound on the maximal value of the function $Z(R)$, to be defined shortly, and $g_2(r)$. The function $Z({\bf r}_i,R)$ is defined for packings of nonoverlapping spheres of unit diameter as the number of sphere centers that are within distance $R$ from a (additional) sphere center at position ${\bf r}_i$, with $i$ an index over centers. The maximum over all ${\bf r}_i$ of $Z({\bf r}_i,R)$ is an upper bound on the maximum of the (different) function $Z(R)$, where $Z(R)$ is defined for a statistically homogeneous packing as the expected number of sphere centers within distance $R$ from any given sphere center, or equivalently as the average of $Z({\bf r}_i,R)$ over all $i$. The function $Z(R)$ can be related to the pair correlation function $g_2(r)$, where for a pair correlation function $g_2({\bf r})$ that is direction-dependent, $g_2(r)$ is the directional average of $g_2({\bf r})$, by \begin{equation} Z(R) = \rho s_1(1)\int_0^Rx^{d-1}g_2(x)dx. \label{ZR} \end{equation} In Eq. (\ref{ZR}), $\rho$ is the constant number density of sphere centers and $s_1(r)$ is the surface area of a sphere of radius $r$ in ${\mathbb R}^d$, \begin{equation} s_1(r) = \frac{2\pi^{d/2}r^{d-1}}{\Gamma(d/2)}. \label{s1} \end{equation} The maximum at fixed $R$ of the function $Z({\bf r}_i,R)$ over all possible configurations of sphere centers $\{{\bf r}_i\}$ is equal to the greatest number $N$ for which $R_{min}(N) \leq R$ for a DLP optimal packing of $N$ spheres in ${\mathbb R}^d$. Defining $Z_{max}(R)$ for all $R$ as this greatest $N$, it follows that \begin{equation} Z(R) \leq Z_{max}(R) \label{ZRbound} \end{equation} for any sphere packing. Equation (\ref{ZRbound}) is a realizability condition on $g_2(r)$, with $Z(R)$ in ${\mathbb R}^d$ defined in terms of $g_2(r)$ in Eq. (\ref{ZR}) and $Z_{max}(R)$ defined completely by the solutions to the DLP problem over all $N$. In ${\mathbb R}^2$, the function $Z_{max}(R)$ may be compared to the function $Z_{tri}(R)$, with $Z_{tri}(R)$ defined as the sum of the number of disk centers included in all (full) shells of radius less than or equal to $R$ in a triangular lattice of contacting disks. Both $Z_{max}(R)$ and $Z_{tri}(R)$ increase roughly linearly with $R^2$, as the area of a disk of radius $R$ is proportional to $R^2$. Clearly, $Z_{max}(R) \geq Z_{tri}(R)$ for all $R$, as can be seen in Fig. \ref{ZmaxVStri}, a plot of $Z_{max}(R)$ vs. $R^2$ for $N=1\!-\!109$ and $d=2$ alongside a plot of $Z_{tri}(R)$. \begin{figure}[ht] \centering \includegraphics[width = 5.5in,viewport = 20 25 735 540,clip]{fig2} \caption{$Z_{max}(R)$ vs $R^2$, as determined by optimal and putative optimal solutions to the DLP problem for $N=1$ to $N=109$, and $Z_{tri}(R)$. The radius $R$ of the (larger) disk enclosing the centers of the $N$ (smaller) disks and fixed disk is measured in units of the diameter of the enclosed disks.} \label{ZmaxVStri} \end{figure} In arbitrary dimension $d$, the function $Z_{max}(R)$ is zero for $R < 1$ due to the nonoverlap condition. For $R=1$, $Z_{max}(R)$ in ${\mathbb R}^d$ is equal to the kissing number $K_d$. For $R > 1$, $Z_{max}(R)$ should grow approximately as $R^d$ in proportion with the growth of the volume of a $d$-dimensional sphere, though in a separate work \cite{HST2009b} we have proved that this cannot be the case for $R \leq \tau$, with $\tau = (1+\sqrt(5))/2$ the golden ratio. For $R \leq \tau$, $Z_{max}(R)$ in any dimension cannot exceed the maximal number of sphere centers that can be placed on the surface of a sphere of radius $R$. Alternatively stated, this counterintuitive result requires that for $R \leq \tau$, $Z_{max}(R)$ can grow only as the surface area $R^{d-1}$. Specifically for $d=2$, $3$ and $4$, $Z_{max}(R\leq\tau)$ is less than or equal to $10$, $33$ and $120$, respectively. \section{Bounds on infinite sphere packings and the DLP problem} We discuss two distinct methods through which the function $Z_{max}(R)$ in ${\mathbb R}^d$ can be employed to bound from above the maximal infinite-volume packing fraction $\phi_*^{\infty}$ of an infinite packing of identical nonoverlapping spheres. The first has been discussed in detail in two separate works \cite{HST2009a,HST2009b}; it is precisely the method of Cohn and Elkies in \cite{CE2003a}. In Ref. \cite{CE2003a}, the authors employ an infinite-dimensional linear program that is the dual of the $g_2$-invariant program \cite{TS2006a} discussed in Sec. I to find the best known bounds on the maximal infinite-volume packing fraction $\phi_*^{\infty}$ of sphere packings at least in dimensions four through 36. An improved method to bound $\phi_*^{\infty}$ from above adds the information encoded in the $Z_{max}(R)$ realizability condition to augment the approach of Cohn and Elkies in \cite{CE2003a} as proposed by Cohn, Kumar and Torquato \cite{CKT2009a}. The second method bounds $\phi_*^{\infty}$ from above by the maximal local packing fraction $\hat{\phi}_*(N)$ of a packing of a number $N$ of identical nonoverlapping spheres around an additional central sphere. The local packing fraction $\hat{\phi}(N)$, of which $\hat{\phi}_*(N)$ is the maximum, is defined for $N$ spheres around an additional fixed central sphere as the total volume of the $N+1$ spheres divided by the volume of a sphere of radius $R$, where $R$ is, as in the DLP problem, the greatest of the distances from the centers of the $N$ surrounding spheres to the center of the fixed sphere. From this definition of $\hat{\phi}(N)$, the maximal local packing fraction $\hat{\phi}_*(N)$ for $N$ $d$-dimensional spheres of unit diameter takes the form \begin{equation} \hat{\phi}_*(N) = \frac{N+1}{(2R_{min}(N))^d}, \label{maxLocalDensity} \end{equation} where $R_{min}(N)$ is as before the optimal radius in the DLP problem for $N$ spheres in ${\mathbb R}^d$. For the sake of convenience, we have collected and defined the various packing fraction terms used in this section in table \ref{packTable}. \begin{table} \caption{Packing fraction terms and definitions used in the text.} \label{packTable} \begin{center} \begin{tabular}{|c c l|} \hline {\bf Symbol} & {\bf Term} & {\bf Definition} \\ \hline \hline \multirow{2}{*}{$\phi(R)$} & \multirow{2}{*}{\,\,\,{\bf packing fraction}\,\,\,} & volume fraction of a (larger) sphere of radius $R$ covered \\ & & by identical nonoverlapping spheres of unit diameter \\ \hline \multirow{2}{*}{$\phi_{max}(R)$} & {\bf maximal} & \multirow{2}{*}{greatest achievable $\phi(R)$ for a given $R$} \\ & {\bf packing fraction} & \\ \hline \multirow{2}{*}{$\phi^{\infty}$} & {\bf infinite-volume} & fraction of space covered by identical nonoverlapping \\ & {\bf packing fraction} & spheres in a given infinite packing \\ \hline \multirow{3}{*}{$\phi_*^{\infty}$} & {\bf maximal} & \\ & {\bf infinite-volume} & greatest achievable infinite-volume packing fraction \\ & {\bf packing fraction} & \\ \hline \multirow{4}{*}{$\hat{\phi}(N)$} & & ratio of the sum of the volumes of a nonoverlapping fixed \\ & {\bf local} & central sphere of unit diameter and its surrounding $N$ \\ & {\bf packing fraction} & same-size spheres to the volume of a (larger) sphere of \\ & & radius $R$, with $R$ defined as in the DLP problem \\ \hline \multirow{2}{*}{$\hat{\phi}_*(N)$} & {\bf maximal local} & \multirow{2}{*}{greatest achievable $\hat{\phi}(N)$ for a given $N$ in ${\mathbb R}^d$; see Eq. (\ref{maxLocalDensity})\,\,} \\ & {\bf packing fraction} & \\ \hline \multirow{2}{*}{$\bar{\phi}(R)$} & {\bf average local} & average packing fraction within a window of radius $R$ of \\ & {\bf packing fraction} & identical nonoverlapping spheres of unit diameter \\ \hline \end{tabular} \end{center} \end{table} The statement that $\hat{\phi}_*(N)$ bounds from above $\phi_*^{\infty}$ for certain $N$ relies on a construction that links local packing fraction $\hat{\phi}(N)$ to the infinite-volume packing fraction $\phi^{\infty}$ of a packing of identical nonoverlapping spheres. The construction proceeds as follows. First, a spherical window of radius $R_{min}(N)$ is centered on an arbitrary sphere in a single configuration of an infinite packing of identical nonoverlapping spheres of unit diameter. An infinite packing of identical nonoverlapping spherical windows is created by replicating the initial window infinitely many times and placing the (replicated) window centers in the exact (scaled) spatial configuration of the centers of the original infinite packing of spheres of unit diameter. The only difference between the two configurations is that the configuration of windows is scaled by $2R_{min}(N)$, the ratio of the radius of a window to the radius of a sphere of unit diameter. As will be made precise in the following paragraphs, for any such packing of windows and spheres of unit diameter, a rigid rotation for the overlayed packing of windows can be found such that the average local packing fraction $\bar{\phi}(R_{min}(N))$ of spheres of unit diameter within the windows is equal to the infinite-volume packing fraction $\phi^{\infty}$ of the spheres of unit diameter in ${\mathbb R}^d$. The concepts of overlay and rotation are illustrated in Fig. \ref{spheresWindows} for a triangular lattice of disks of unit diameter. \begin{figure}[ht] \centering \includegraphics[width = 5.8in,viewport = 0 120 815 425,clip]{fig3} \caption{Illustration of a rotation of an infinite packing of identical nonoverlapping windows of radius $R$, arranged on the sites of a triangular lattice, overlayed upon an infinite packing of smaller identical nonoverlapping disks. A rotation is selected such that irrational ratios are achieved between the components of at least one of the lattice vectors of the packing of windows in the directions of the lattice vectors of the packing of smaller disks. As a result, at large distances from the axis of rotation, any window can be thought of as being placed at random onto the packing of smaller disks. It follows that the average local packing fraction $\bar{\phi}(R)$ of the smaller disks within the windows is equal to the infinite-volume packing fraction $\phi^{\infty}$ of the smaller disks.} \label{spheresWindows} \end{figure} It suffices to apply the aforementioned construction to periodic packings, as it has been shown that periodic packings in ${\mathbb R}^d$ can obtain an infinite-volume packing fraction $\phi^{\infty}$ arbitrarily close to the maximal infinite-volume packing fraction $\phi_*^{\infty}$ (for example, see \cite{CE2003a}). A periodic packing can be defined in terms of a lattice $\Lambda$, where $\Lambda$ in ${\mathbb R}^d$ is a subgroup consisting of the integer linear combinations of a set of vectors that constitute a basis for ${\mathbb R}^d$. For identical nonoverlapping spheres, a lattice packing is a packing where the centers of the spheres are located at the points of $\Lambda$. In such a lattice packing, the space ${\mathbb R}^d$ can be divided into finite-size identical nonoverlapping regions called fundamental cells, each containing the center of only one sphere. A periodic packing is a more general formulation of a lattice packing. For identical nonoverlapping spheres, a periodic packing is obtained by placing a fixed configuration of a number $M$ of spheres in a fundamental cell that is then periodically replicated (without overlap between cells or spheres) to cover ${\mathbb R}^d$. The fixed configuration of $M$ spheres within each fundamental cell is arbitrary subject only to the overall nonoverlap condition of the periodic packing of spheres. As used here, the term ``lattice'' is the same as ``Bravais lattice'' conventionally used in the physics literature. Consider an infinite periodic packing (with lattice basis vectors $\{{\bf u}_j\}$) of nonoverlapping spheres of unit diameter in ${\mathbb R}^d$, $d > 1$, with infinite-volume packing fraction $\phi^{\infty}$. Place an infinite periodic packing (with lattice basis vectors $\{{\bf v}_j\}$) of identical nonoverlapping windows of radius $R_{min}(N)$ over the infinite periodic packing of spheres of unit diameter in the manner of the construction discussed previously. For the radius $R_{min}(N)$ of the windows, any positive integer $N_* \in {\mathbb N}$ can be considered, where in ${\mathbb R}^d$ the set ${\mathbb N}$ is defined such that each $N_*$ is the greatest number $N$ of spheres for which $R_{min}(N_*) = R_{min}(N)$. For example, with $N = 1$ in two dimensions, $R_{min}(1) = 1$, and the greatest $N$ for which $R_{min}(N) = 1$ is $N=N_*=6$ \cite{endnote5}. It is intuitively clear that for any greatest such $N = N_*$ and $R = R_{min}(N_*)$ that \begin{equation} \hat{\phi}_*(N_*) \geq \phi_{max}(R_{min}(N_*)), \label{phiStarVsPhiMax} \end{equation} where $\phi_{max}(R)$ is defined as the maximal fraction of space that identical nonoverlapping spheres of unit diameter may cover in a spherical window of radius $R$ \cite{endnote6}. Now rigidly rotate the packing of windows around the center of the original window as in Fig. \ref{spheresWindows}. The component of a window-packing basis vector ${\bf v}_n$ that is in the direction of a given unit-diameter sphere packing basis vector ${\bf u}_m$ is ${\bf v}_n \cdot {\bf u}_m/|{\bf u}_m|$. A rotation may be found such that the ratios $\{{\bf v}_n \cdot {\bf u}_j/|{\bf u}_j|^2\}$ of the components of (at least) one of the window basis vectors ${\bf v}_n$ in the directions of each of the sphere packing basis vectors $\{{\bf u}_j\}$ to the respective magnitudes $\{|{\bf u}_j|\}$ of the sphere packing basis vectors are all irrational. This concept is illustrated in Fig. \ref{cellsWindows} in ${\mathbb R}^2$ for parallelogram fundamental cells. After the rotation, due to the irrationality of the ratios of lattice vector components, the average fraction of space $\bar{\phi}(R_{min}(N))$ covered by the spheres of unit diameter in each window of the window packing is equal to the infinite-volume packing fraction $\phi^{\infty}$ of the unit-diameter spheres. \begin{figure}[ht] \centering \includegraphics[width = 5.8in,viewport = 5 190 715 430,clip]{fig4} \caption{Illustration of a rigid rotation of a window packing with parallelogram fundamental cells such that the ratios of the components of ${\bf v}_1$ in the directions of each of ${\bf u}_1$ and ${\bf u}_2$ to the magnitudes of ${\bf u}_1$ and ${\bf u}_2$, respectively, are both irrational.} \label{cellsWindows} \end{figure} \begin{figure}[ht] \centering \includegraphics[width = 5.5in,viewport = 20 35 765 525,clip]{fig5} \caption{maximal packing fraction $\phi_*^{\infty} = 0.906900\dots$ compared to putative upper bounds on $\phi_*^{\infty}$ for $d=2$ as determined by putative optimal solutions to the DLP problem for selected $N_* \in {\mathbb N}$ up to $N_*=348$. The minimum upper bound determined here is at $N_*=336$, $\hat{\phi}_*(336) = 0.928114$.} \label{upperBoundPlot} \end{figure} With this construction, the packing fraction $\phi^{\infty}$ of the spheres of unit diameter, equal to the average local packing fraction $\bar{\phi}(R_{min}(N))$ of the unit-diameter spheres within a window, can be compared to the maximal local packing fraction $\hat{\phi}_*(N)$ of that same window. As $\phi_{max}(R_{min}(N))$ is clearly greater than or equal to $\phi^{\infty}$, using Eq. (\ref{phiStarVsPhiMax}) we have \begin{equation} \hat{\phi}_*(N_*) \geq \phi_{max}(R_{min}(N_*)) \geq \phi^{\infty}, \label{upperBoundInt} \end{equation} for $N_* \in {\mathbb N}$ in ${\mathbb R}^d$. Equation (\ref{upperBoundInt}) is true for all feasible $\phi^{\infty}$ including the maximal infinite-volume packing fraction $\phi_*^{\infty}$ of nonoverlapping spheres in ${\mathbb R}^d$. The maximal local packing fraction $\hat{\phi}_*(N_*)$ is therefore an upper bound on the maximal infinite-volume packing fraction $\phi_*^{\infty}$, or \begin{equation} \hat{\phi}_*(N_*) \geq \phi_*^{\infty}, \qquad N_* \in {\mathbb N}. \label{upperBound} \end{equation} Figure \ref{upperBoundPlot} plots for $d=2$ and for selected values of $N_* \in {\mathbb N}$ up to $N_*=348$ the putative upper bound defined by $\hat{\phi}_*(N_*)$ alongside the proved maximal packing fraction, $\phi_*^{\infty} = \pi / \sqrt{12}$. While the maximal infinite-volume packing fractions for $d$-dimensional identical nonoverlapping spheres are known with analytical rigor for $d=1\!-\!3$, $d=8$ and $d=24$ \cite{CK2010a}, this method could be used to improve upon the upper bounds on maximal packing fraction $\phi_*^{\infty}$ in dimensions where a value for $\phi_*^{\infty}$ has not not yet been proved. It is important to reiterate though that to be rigorous, the upper bound as defined requires a DLP radius $R_{min}(N)$ that is proved optimal. \section{Optimal packings in two dimensions} The packing of nonoverlapping disks that uniquely achieves the highest infinite-volume packing fraction $\phi_*^{\infty} = \pi/\sqrt{12}$ for $d=2$ is well-known to be the packing such that each disk is in contact with exactly six others, with centers arranged on the sites of the triangular lattice. As DLP optimal packings are packing fraction-maximizing arrangements of $N$ disks centered on a (additional) disk, one might expect that all or a major subset of disks in any given DLP optimal packing will always sit on the sites of the triangular lattice. Over the range of $N$ studied, however, this is in fact very infrequently the case. We find that at only three values of $N$ greater than $6$, at $N = 12$, $30$ and $54$, are triangular lattice configurations also DLP optimal configurations, while most DLP optimal packings are significantly more locally dense than a packing of $N$ disks around a central disk with centers on the sites of the triangular lattice, as is illustrated in Fig. \ref{ZmaxVStri} for $N=1\!-\!109$. In general, we find wide variation in the symmetries and other characteristics of DLP optimal packings. For the majority of cases, there appear to be an uncountably infinite number (a continuum) of optimal configurations of $N$ sphere centers at optimal $R_{min}(N)$, with the continuum attributable to the presence of rattlers. A rattler in a packing of spheres in ${\mathbb R}^d$ is a sphere that is positioned such that it may be individually moved in at least one direction without resulting overlap of any other sphere within the packing or the packing boundary (in this case, the encompassing sphere of radius $R_{min}(N)+0.5$), i.e., a rattler is a sphere that is not locally jammed \cite{TTD2000a,TS2001a}. The rattlers present in the following figures are indicated by a lighter shading, while the fixed central spheres (disks) are indicated by an open circle. \subsection{Packings that are proved optimal} In two dimensions for $N \leq 10$, we proved \cite{HST2009b} that optimal packings are those in which $R_{min}(N)$ is equal to the radius of the minimal-radius circle onto which the centers of $N$ disks may be packed. These $R_{min}(N)$ may be analytically calculated via simple trigonometry, yielding $R_{min}(N) = 1$ for $N \leq 6$, and \begin{equation} R_{min}(N) = \frac{1}{\sqrt{2(1-\cos{2\pi/N})}}, \qquad 6 \leq N \leq 10. \label{Rmin10} \end{equation} For $6 \leq N \leq 9$, DLP optimal packings with $R_{min}(N)$ as defined in Eq. (\ref{Rmin10}) are unique up to rotations and correspond to configurations where all sphere centers lie on the circle of radius $R_{min}(N)$ at distance unity from two adjacent centers. In general, for many $N$ there is a unique (up to rotations) optimal packing, and in a smaller number of cases such as for $N=10$, $60$, $90$ and $126$, there are a finite number of degenerate optimal packings. Figure \ref{10opt} depicts three of a finite number of optimal packings for $N=10$, $d=2$, where $R_{min}(10) = \tau$; these packings are formed by radially translating up to five disks with centers on the circle at radius $R_{min}(10) = \tau$, the golden ratio, to be in contact with the fixed central disk at distance unity. \begin{figure}[!ht] \centering \subfigure[]{\includegraphics[width=2.0in]{fig6a}} \\ \subfigure[]{\includegraphics[width=2.0in]{fig6b}} \\ \subfigure[]{\includegraphics[width=2.0in]{fig6c}} \caption{Three of a finite number of optimal cases for $N=10$, $R_{min}(10) = \tau = 1.618204\dots$ formed by radially translating disks whose centers lie on a circle of radius $R_{min}(10) = \tau$ to be in contact with the fixed central disk. (a) point group $C_{2v}$. (b) point group $C_{2v}$. (c) point group $D_{5h}$.} \label{10opt} \end{figure} \subsection{Curved hexagonal packings} For $N > 10$, we know of no packings that have been proved optimal. However, for certain $N$, previous studies have found packings that we conjecture to be DLP optimal packings. For $N+1$ equal to hexagonal number $3k(k+1)+1$, $k \geq 1$ an integer, Lubachevsky and Graham \cite{LG1997a} found a class of packings called \textit{curved hexagonal} packings that they conjectured to be the densest packings of $N+1$ identical nonoverlapping disks within an encompassing disk for $k=1\dots5$. The characteristics of this class include that each packing has a disk fixed at the center of the encompassing disk, and that all curved hexagonal packings belong to point group $C_{6h}$, meaning that they are invariant under $60^{\circ}$ rotation or inversion through the origin. Further, for $k \geq 4$, for each $k$ there are a finite number of degenerate packings of equal density (for $1 \leq k \leq 3$, there is a unique densest packing). The degenerate packings are chiral; each ring of disks, or disks that share a common radius, beginning with the fourth ring (the fourth farthest from the center), can be angularly oriented in more than one distinct fashion relative to the preceding ring. By reorienting rings, the degenerate packings for a given $k$ can be generated from one another. Figure \ref{curvedHex} depicts curved hexagonal packings for $N=60$, $90$ and $126$. \begin{figure}[!ht] \centering \subfigure[]{\includegraphics[width=2.0in]{fig7a}} \\ \subfigure[]{\includegraphics[width=2.0in]{fig7b}} \\ \subfigure[]{\includegraphics[width=2.0in]{fig7c}} \caption{Curved hexagonal packings that are also DLP optimal packings. (a) $N=60$, $R_{min}(60) = 3.830649\dots$, point group $C_{6h}$. (b) $N=90$, $R_{min}(90) = 4.783386\dots$, point group $C_{6h}$. (c) $N=126$, $R_{min}(126) = 5.736857\dots$, point group $C_{6h}$.} \label{curvedHex} \end{figure} DLP optimal packings for $N$ disks are equivalent to the densest packings of $N+1$ disks enclosed in an encompassing disk when one of the disks is fixed at the center. For $N=6$, $18$, $36$, $60$ and $90$ ($k=1\dots5$), we find that the curved hexagonal packings are the only DLP optimal packings, in support of the conjecture of Lubachevsky and Graham that curved hexagonal packings are the densest packings up to $k=5$. Further, for $N=126$ ($k = 6$), we also find that curved hexagonal packings are the only DLP optimal packings, indicating that though there are packings of $127$ unconstrained disks within an encompassing disk denser than the curved hexagonal packings (as were found by Lubachevsky and Graham \cite{LG1997a}), the curved hexagonal packings remain the densest packings of $126$ disks around a fixed central disk. This is not the case for $N=168$ ($k = 7$), as we find DLP optimal packings with higher density, such as the $N=168$ packing to be shown later in the top panel of Fig. \ref{bigDiff}. \subsection{Wedge hexagonal packings} Another class of packings, previously unidentified, contain a subset of disks with centers arranged on the sites of the triangular lattice and the remainder arranged in six ``wedges''. We hereafter term such packings \textit{wedge hexagonal} packings. Wedge hexagonal packings are not DLP optimal packings when arranged symmetrically (point group $D_{6h}$); however, minor deviations from perfect symmetry in a wedge hexagonal packing can produce a DLP optimal packing. Figure \ref{wedgeHexA} depicts such DLP optimal packings for $N=84$, $120$ and $162$. Lines to guide the eye have been drawn on the three optimal packings in Fig. \ref{wedgeHexA}. \begin{figure}[!ht] \centering \subfigure[]{\includegraphics[width=2.0in]{fig8a}} \\ \subfigure[]{\includegraphics[width=2.0in]{fig8b}} \\ \subfigure[]{\includegraphics[width=2.0in]{fig8c}} \caption{DLP optimal packings that are minor deviations from ``wedge hexagonal'' packings. (a) $N=84$, $R_{min}(84) = 4.581556\dots$, $(k,p,a) = (3,2,1)$, point group $C_i$. (b) $N=120$, $R_{min}(120) = 5.562401\dots$, $(k,p,a) = (3,3,1)$, point group $C_i$. (c) $N=162$, $R_{min}(162) = 6.539939\dots$, $(k,p,a) = (3,4,1)$, point group $C_i$.} \label{wedgeHexA} \end{figure} In a wedge hexagonal packing, the subset of disks with centers arranged on the sites of the triangular lattice contains two parts; a regular hexagonal core of hexagonal number $3k(k+1) + 1$ disks, with $k \geq 3$ odd; and six `branches' composed of $(pk-a)$ disks, with $p \geq 2$ and $a \geq 1$ integers, extending from each of the vertices of the core regular hexagon. The branches are $k$ disks wide and $p$ disks long, with $a$ of the farthest disks removed such that the end of the branch approximates a circle (as opposed to the point of a triangle). The six wedges are arranged roughly as $p(p+1)/2$ bowling pins and lie in between the branches, with each of the six `lead pins' placed at the midpoint of each side of the core hexagon. The DLP optimal packings in Fig. \ref{wedgeHexA} are the wedge hexagonal packings, with minor deviations in the positions of some disks, for $(k,p,a) = (3,2\!-\!4,1)$. In general, the deviations necessary to produce a DLP optimal packings from a wedge hexagonal packing occur in the branches, and to a lesser degree, the wedges of the packing, while the core regular hexagon retains perfect six-fold symmetry. The deviations required differ for each wedge hexagonal packing, but from our observations produce packings where the backbone maintains inversion symmetry through the origin, where the backbone of a packing is defined as the packing excluding the rattlers. Such deviations can be seen in the branches and wedges in the DLP optimal packings for $N=198$ and $N=312$ in Fig. \ref{wedgeHexB}, which correspond to the (slightly altered) wedge hexagonal packings with $(k,p,a) = (5,3,3)$ and $(7,3,3)$, respectively. \begin{figure}[!ht] \centering \subfigure[]{\includegraphics[width=2.0in]{fig9a}} \\ \subfigure[]{\includegraphics[width=2.0in]{fig9b}} \caption{DLP optimal packings that are minor deviations from the ``wedge hexagonal'' packings. (a) $N=198$, $d=2$. $R_{min}(198) = 7.201130\dots$, $(k,p,a) = (5,3,3)$, point group $C_{i}$. (b) $N=312$, $R_{min}(312) = 9.141107\dots$, $(k,p,a) = (7,3,3)$, point group $C_i$.} \label{wedgeHexB} \end{figure} \subsection{DLP optimal packings with high symmetry} Many DLP optimal packings exhibit symmetries other than inversion symmetry through the origin as exhibited by the altered wedge hexagonal packings shown in Figs. \ref{wedgeHexA} and \ref{wedgeHexB}. These symmetries include perfect bond orientational order, invariance under rotation through an angle, and invariance under reflection across an axis. A list of packing point group, alongside other packing properties such as $R_{min}(N)$ value, of all of the DLP optimal packings depicted in this work appears in Appendix \ref{tables}. \begin{figure}[!ht] \centering \subfigure[]{\includegraphics[width=2.0in, viewport = 30 95 595 680,clip]{fig10a}} \\ \subfigure[]{\includegraphics[width=2.0in, viewport = 30 95 595 680,clip]{fig10b}} \\ \subfigure[]{\includegraphics[width=2.0in, viewport = 30 95 595 680,clip]{fig10c}} \caption{Three examples of DLP optimal packings incorporating interesting symmetry elements. (a) $N=25$, $R_{min}(25) = 2.497212\dots$, point group $D_{5h}$. (b) $N=11$, $R_{min}(11) = 1.685854\dots$, point group $C_{2v}$. (c) $N=32$, $R_{min}(32) = 2.794164\dots$, point group $D_{2h}$.} \label{oddSyms} \end{figure} Perfect five-fold symmetry, disallowed to regular infinite crystals, is exhibited by three of the optimal packings studied. Five-fold rotational symmetry is present in the $N=15$ (Fig. \ref{DLP15pic}), $N=10$ (bottom panel of Fig. \ref{10opt}) and $N=25$ (top panel of Fig. \ref{oddSyms}) packings. The $N=25$ optimal packing also has perfect five-fold bond orientational order, evident in that all nearest-neighbor disk pairs are at one of five angles relative to any fixed coordinate system. Additionally, it is of note that the $N=25$ packing may be tiled by $15$ identical rhombuses of acute angle $72^{\circ}$ with vertices placed at disk centers, where the rhombus of acute angle $72^{\circ}$ is known to be the `thicker' of the two types of rhombus present in a Penrose tiling \cite{Penrose1974a}. The top panel of Fig. \ref{jmolPics}, a diagram of the contact network for the $N=25$ optimal packing, depicts these rhombuses. Two other optimal packings incorporating interesting symmetry elements are the $N=11$ and $N=32$ packings (center and bottom panels of Fig. \ref{oddSyms}, respectively). The packing depicted for $N=11$ belongs to symmetry group $C_{2v}$ and appears to be the unique DLP optimal packing of $11$ disks. The packing for $N=32$ is one of an infinity of possible packings due to the presence of two rattlers; however, the backbone of the packing has reflection symmetry across two axes and inversion symmetry through the origin and belongs to symmetry group $D_{2h}$. Though five-fold symmetry may be limited to packings with small $N$, high symmetry in general is not. For example, the backbone of the optimal packing with the largest $N$ presented here, $N=348$, has six-fold rotation symmetry and belongs to point group $D_{6h}$. Figure \ref{manySpheres} depicts the $N=348$ optimal packing, which contains $24$ rattlers. \begin{figure}[!ht] \centering \includegraphics[width=2.0in]{fig11} \caption{DLP optimal packing for $N=348$, $R_{min}(348) = 9.620709\dots$, belonging to point group $D_{6h}$.} \label{manySpheres} \end{figure} \subsection{Unusual features in select optimal packings} \begin{figure}[!ht] \centering \subfigure[]{\includegraphics[width=2.0in, viewport = 30 95 595 680,clip]{fig12a}} \\ \subfigure[]{\includegraphics[width=2.0in, viewport = 30 95 595 680,clip]{fig12b}} \\ \subfigure[]{\includegraphics[width=2.0in, viewport = 30 95 595 680,clip]{fig12c}} \caption{Three examples of DLP optimal packings with circular cavities about the fixed central disk. (a) $N=24$, $R_{min}(24) = 2.425256\dots$, point group $D_{3h}$. (b) $N=45$, $R_{min}(45) = 3.374023$, point group $C_1$. (c) $N=95$, $R_{min}(95) = 4.958096$, point group $C_1$.} \label{circCavities} \end{figure} A prevalent feature found in many of the DLP optimal packings studied is a cavity consisting of a ring of disks enclosing, but not contacting, the fixed central disk. The counterintuitive presence of this feature is related to the aforementioned fact that in any dimension $d$, no spherical window of radius $R \leq \tau$ centered on a central nonoverlapping sphere of unit diameter may encircle more sphere centers than the number (plus one) that can be placed on the encircling sphere's surface. Circular cavities around the fixed central disk appear in many DLP optimal packings, including those for $N=24$, $45$ and $95$ disks, as illustrated in Fig. \ref{circCavities}. There are $9$, $8$ and $7$ disks, respectively, forming the walls of the cavities in the three packings in Fig. \ref{circCavities}. Two particularly notable DLP optimal packings that include a cavity enclosing the fixed central disk are the $N=40$ and $N=66$ packings shown in Fig. \ref{eyePackings}. \begin{figure}[ht] \centering \subfigure[]{\includegraphics[width=2.0in]{fig13a}} \\ \subfigure[]{\includegraphics[width=2.0in]{fig13b}} \caption{DLP optimal packings consisting of layers of distorted rings, or `eyes', enclosing the central disk. (a) $N=40$, $R_{min}(40) = 3.136712\dots$, point group $D_{2h}$. (b) $N=66$, $R_{min}(66) = 4.104997$, point group $C_1$.} \label{eyePackings} \end{figure} These packings are composed of layers of distorted rings, where the distorted rings appear as eye-like closed curves of varying curvature with each successive layer from the center more circular than the last. It is curious that even though the only shape imposed upon the packings, in the form of the encompassing disk and the disks themselves, is circular, that an optimal packing incorporating distorted rings emerges for these numbers of disks. The bottom panel of Fig. \ref{jmolPics} is a diagram of the contact network for the backbone of the $N=40$ optimal packing. \begin{figure}[!ht] \centering \subfigure[]{\includegraphics[width=2.0in, viewport = 30 125 595 635,clip]{fig14a}} \\ \subfigure[]{\includegraphics[width=2.0in, viewport = 30 125 595 635,clip]{fig14b}} \caption{Diagrams \cite{jmolRef} of contact networks for the $N=25$ (a) and $N=40$ (b) optimal packings, with point group symmetries $D_{5h}$ and $D_{2h}$, respectively.} \label{jmolPics} \end{figure} The presence of these cavities about the central disk leads to an interesting counterintuitive result. Suppose that in a binary liquid of nonoverlapping disks of unit diameter, one species of disk is endowed with an attractive square well potential. The potential of the ``attractive'' disks, present in the dilute limit in comparison to the ``nonattractive'' disks, acts on the centers of the ``nonattractive'' disks extending only to a distance just larger than any of the $R_{min}(N)$ presented in Figs. \ref{circCavities} or \ref{eyePackings}. If the depth (strength) of this square well were made arbitrarily large, then the result would seem paradoxical: unbounded attraction to the disks, in a minimal energy configuration, would eliminate contact with these disks. This effect in such a binary liquid of disks requires that the pair correlation function depicting the probability density of finding the centers of a given number of ``nonattractive'' disks a certain distance from the centers of ``attractive'' disks be zero from $r=0$ to $r = R_0 > 1$ a specified distance in excess of the diameter of the disks. None of the currently available pair correlation function theories are able to predict this effect, including the crucial dependence of $R_0$ on $R_{min}(N)$, because the underlying approximations of the currently available theories cannot account for the basic many-body geometrical features involved. \subsection{Imperfect symmetry} \begin{figure}[!ht] \centering \subfigure[]{\includegraphics[width=2.0in, viewport = 30 95 595 680,clip]{fig15a}} \\ \subfigure[]{\includegraphics[width=2.0in, viewport = 30 95 595 680,clip]{fig15b}} \\ \subfigure[]{\includegraphics[width=2.0in, viewport = 30 95 595 680,clip]{fig15c}} \caption{DLP optimal packings that exhibit imperfect symmetry. (a) $N=59$, $R_{min}(59) = 4.824374$, point group $C_1$. (b) $N=80$, $R_{min}(80) = 4.514170$, point group $C_1$. (c) $N=46$, $R_{min}(46) = 3.414304$, point group $C_1$.} \label{noSym} \end{figure} Not all DLP optimal packings exhibit perfect symmetry; for many $N$, a subset of disks in an optimal configuration appear to mimic a symmetric packing, but the packing as a whole exhibits only imperfect symmetry. One situation in which this occurs frequently is when the number of disks $N$ in the optimal packing is close to a different number for which the optimal packing is relatively unusually dense. For example, the optimal $N=59$ packing shown in the top panel of of Fig. \ref{noSym} lacks any of the symmetry elements described above but nonetheless closely resembles the particularly dense $N=60$ curved hexagonal packing (top panel of Fig. \ref{curvedHex}). Other packings in which imperfect symmetry is present include the $N=80$ packing shown in the center panel of Fig. \ref{noSym}, the disks closer to the center of which are ordered with centers on the sites of a slightly distorted triangular lattice, and the $N=46$ packing shown in the bottom panel of Fig. \ref{noSym}, which has imperfect five-fold symmetry. The $N=46$ packing, along with the $N=45$ packing (center panel of Fig. \ref{circCavities}), together illustrate another finding: that the structure of optimal packings even for consecutive numbers of disks can vary substantially. \subsection{Surface effects} DLP optimal packings with $N$ in the higher range of the packings studied appear, as $N$ increases, to more and more resemble the triangular lattice in the bulk of the packing. Nonetheless, the surface of the packing always deviates significantly from the bulk crystal. In general, the optimal packings at higher $N$ consist of a ``bulk zone'' with disk centers arranged in the triangular lattice surrounded by a ``surface zone'' with disk centers arranged in circular rings. This effect can be seen in all of the optimal packings with $N$ in the higher range of $N$ studied, including in those shown in Figs. \ref{wedgeHexA}, \ref{wedgeHexB}, \ref{manySpheres}, in the center panel of Fig. \ref{noSym} and in Fig. \ref{bigDiff}. Qualitatively, it appears that the radial width of the surface zone tends to increase with number of disks, though not as fast as the radial width of the bulk. Due to computational time constraints, we were unfortunately not able to quantitatively verify this result at much larger $N$; however, should the observed trend continue, the width of the surface zone would continue to grow as the bulk grows, eventually becoming infinitely large as $N\rightarrow\infty$. This does not imply that at very large $N$ the surface zone would represent a substantial fraction of the total packing; as $N\rightarrow\infty$, the ratio of space covered by the surface zone to the space covered by the bulk zone will still be zero. \begin{figure}[!ht] \centering \subfigure[]{\includegraphics[width=2.0in]{fig16a}} \\ \subfigure[]{\includegraphics[width=2.0in]{fig16b}} \\ \subfigure[]{\includegraphics[width=2.0in]{fig16c}} \caption{Three packings for which $R_{min}(N)$ is significantly less than the radius of a disk enclosing the centers of $N+1$ disks arranged on the sites of the triangular lattice. (a) $N=168$, $R_{min}(168) = 6.680013$, point group $C_1$. (b) $N=264$, $R_{min}(264) = 8.417769$, point group $C_1$. (c) $N=270$, $R_{min}(270) = 8.497744$, point group $C_1$.} \label{bigDiff} \end{figure} For certain optimal packings exhibiting this surface effect, we also find that the optimal radius $R_{min}(N)$ is significantly smaller than the radius of the smallest disk centered on a fixed central disk that can enclose (an additional) $N$ disks with centers arranged on the sites of the triangular lattice. Figure \ref{bigDiff} depicts three optimal packings for which this difference is relatively large, at $N=168$, $264$ and $270$, where the $R_{min}(N)$ differ by $0.248190$, $0.128616$ and $0.162510$ from the respective radii for the triangular lattice configuration, $7$, $\sqrt{73}$ and $\sqrt{75}$. Each of the packings in Fig. \ref{bigDiff} also displays its own interesting features, including a close resemblance on the left of the image of the $N=168$ packing to the $(3,4,1)$ wedge hexagonal packing, and imperfect three-fold rotational symmetry for the $N=264$ and $N=270$ packings. \section{Conclusions and discussion} The DLP problem is a local packing problem that in certain limits encompasses both the infinite sphere packing and kissing number problems. DLP optimal packings exhibit a wide variety of symmetries, vary significantly from packings with disk centers placed on the sites of a triangular lattice, and often vary significantly with consecutive $N$. Two local packing classes, the curved hexagonal and wedge hexagonal packings, lead to the densest or very dense packings of disks over the range of $N$ studied. The optimal radii $R_{min}(N)$ corresponding to a packing of $N$ spheres in ${\mathbb R}^d$ form a realizability condition on functions that are candidates to be the pair correlation function $g_2(r)$ of a statistically homogeneous and isotropic packing of spheres. This realizability condition incorporates more information than is included in the structure factor and pair correlation function nonnegativity conditions alone. Though the condition discussed here only applies to packings of identical spheres, equivalent $Z_{max}(R)$ functions for packings of differentiated non-spherical objects can be found and corresponding realizability conditions imposed. The function $Z_{max}(R)$ can also be employed in the two ways discussed to bound from above the packing fraction of an infinite sphere packing in dimension $d$. Similarly, $Z_{max}(R)$ for differentiated non-spherical objects can be employed to form upper bounds on corresponding infinite packing maximal packing fractions. Our work has direct application to packaging, particularly to problems involving identical nonoverlapping disks within a circular boundary. Further, as is discussed in Appendix \ref{algorithm}, the algorithm we have employed to find the $d=2$ putative DLP optimal packings presented in this work may be modified to study dense packings of $d$-dimensional differentiated objects of various shape within different boundaries. Additionally, our work has implications for nucleation theory and surface physics, particularly in terms of the effects of imposing a circular boundary upon a packing of spheres with centers initially placed on the sites of a triangular lattice. In future work, we expect to investigate these implications and others in more depth. In a sequel to this paper, we will present and analyze DLP optimal packings and their corresponding $R_{min}(N)$ for three-dimensional spheres over a larger range of $N$. We will catalogue optimal packings of particularly high packing fraction and of unusual symmetry, and we will investigate the possibility of $d=3$ extensions to special $d=2$ classes of packings such as the curved hexagonal and wedge hexagonal packings. We will compare $R_{min}(N)$ values to shell distances in Barlow packings \cite{Barlow1883a}, where the Barlow packings, of which the best-known are face centered cubic (FCC) and hexagonal close packed (HCP) arrangements, all individually achieve the maximal infinite-volume packing fraction for $d=3$, $\phi_*^{\infty} = \pi/\sqrt{18} = 0.740481\dots$. Packing in ${\mathbb R}^3$ is intrinsically more complicated than in ${\mathbb R}^2$. In ${\mathbb R}^3$, there is a wider range of possibilities for contact coordination, particularly for twelve contacting spheres around a central sphere (with $K_3 = 12$) where there is an infinity of possible configurations, as compared to just one in ${\mathbb R}^2$. Additionally, there is a single optimal infinite-volume packing configuration in ${\mathbb R}^2$, i.e., the triangular lattice, whereas there is an infinite number in ${\mathbb R}^3$, i.e., the Barlow packings. Preliminary findings indicate that there is less symmetry (as quantified by point groups) present for the majority of DLP optimal packings in ${\mathbb R}^3$ and more variation with $N$. Further, observations suggest that for the same value of $N$ that there are significantly more locally jammed packing configurations in ${\mathbb R}^3$ with radius $R > R_{min}(N)$ than is the case in ${\mathbb R}^2$. This finding has implications for the dynamics of nucleation occuring in pure supersaturated liquids. If a nucleus in a supersaturated liquid of identical nonoverlapping spheres in ${\mathbb R}^3$ is approximated as a group of $N$ densely-packed spheres with centers within distance $R$ of a central sphere, then there are more packing configurations available to a nucleus of radius $R$, with $R$ confined to a small finite range $R > R_{min}(N)$, in ${\mathbb R}^3$ than in ${\mathbb R}^2$. \bigskip \noindent {\bf ACKNOWLEDGEMENTS:} \medskip S.T. thanks the Institute for Advanced Study for its hospitality during his stay there. This work was supported by the Division of Mathematical Sciences at the National Science Foundation under Award Number DMS-0804431 and by the MRSEC Program of the National Science Foundation under Award Number DMR-0820341.
\section{Introduction} The standard model (SM) has a unique way of incorporating CP violation (CPV) and suppressing flavor changing neutral currents (FCNCs). Till today no deviation from the SM predictions related to quark flavor violation has been observed. Regarding the first two generations, models which do not include some sort of degeneracies or flavor alignment (that is, when new physics contributions are diagonal in the quark mass basis) are bounded to a high energy scale. Moreover, contributions involving only quark doublets cannot be simultaneously aligned with both the down and the up mass bases, hence even alignment theories are constrained by measurements. However, the hierarchy problem is not triggered by the light quarks, but rather by the large top Yukawa, where almost any natural new physics (NP) model consists of an extended top sector. Ironically, the top flavor sector is the least understood one, and at present no model independent bound on its coupling is known to exist. In this work, we formulate a simple basis independent formalism for studying flavor constraints in the quark sector (recent related work about algebraic flavor invariants can be found in~\cite{Feldmann:2009dc,Jenkins:2009dy}). We start with a two generations analysis, where a natural geometric interpretation can be applied. It allows us to straightforwardly reproduce known results~\cite{Blum:2009sk}. We then consider the three generations case, where a dramatic improvement in the measurements related to the top sector is expected at the LHC. The combination of data from the down and the up sectors is used to robustly constrain models including arbitrary mechanisms of alignment. In the absence of Yukawa interactions, the SM quark sector possesses a global $G_{\rm SM}=U(3)_Q\times U(3)_U \times U(3)_D$ flavor symmetry, where $Q$, $U$ and $D$ stand for quark doublets, up and down type quark singlets, respectively. $G_{\rm SM}$ is broken by the Yukawa couplings $Y_u$ and $Y_d$, which transform as $(\mathbf{3}, \bar{\mathbf{3}},1)$ and $(\mathbf{3},1,\bar{\mathbf{3}})$, respectively, under the flavor group. The spurions $Y_u Y_u^\dagger$ and $Y_d Y_d^\dagger$ are then both in the $(\mathbf{8}+1,1,1)$ representation. Since the trace of these matrices does not affect flavor changing processes, it is useful to remove it, and work with $(Y_u Y_u^\dagger)_{\slashed{\mathrm{tr}}}$ and $(Y_d Y_d^\dagger)_{\slashed{\mathrm{tr}}}$, adjoints of $SU(3)_Q$. For simplicity of notation, we denote these objects as \begin{eqnarray} \mathcal{A}_u \equiv (Y_u Y_u^\dagger)_{\slashed{\mathrm{tr}}} \, , \qquad \mathcal{A}_d \equiv (Y_d Y_d^\dagger)_{\slashed{\mathrm{tr}}} \, . \end{eqnarray} \section{Two Generations} Any hermitian traceless $2\times 2$ matrix can be expressed as a linear combination of the Pauli matrices. This combination can be naturally interpreted as a vector in 3D real space, which applies to $\mathcal{A}_d$ and $\mathcal{A}_u$. We can then define a length of such a vector, a scalar product, a cross product and an angle between two vectors, all of which are basis independent: \begin{eqnarray} \label{definitions} \hspace*{-.3cm}&&|\vec{A}|^2 \equiv \frac{1}{2} \mathrm{tr}(A^2), \ \vec{A} \cdot \vec{B} \equiv \frac{1}{2} \mathrm{tr}(A B), \ \vec{A} \times \vec{B} \equiv \frac{i}{2} \left[ B,A \right] , \nonumber \\ \hspace*{-.3cm}&&\cos \theta_{AB} \equiv \frac{\vec{A} \cdot \vec{B}}{|\vec{A}| |\vec{B}|}, \ \sin \theta_{AB} =\frac{\left| \vec{A} \times \vec{B} \right|}{|\vec{A}||\vec{B}|}\, , \end{eqnarray} where the two angle definitions are equivalent. This allows for an intuitive understanding of the flavor and CPV induced by a NP source. Consider a dimension six $SU(2)_L$-invariant operator, involving only quark doublets, \begin{eqnarray} \label{o1} \frac{z_1}{\Lambda_{\rm NP}^2}\, O_1=\frac{1}{\Lambda_{\rm NP}^2} \left[ \overline{Q}_{i} (X_Q)_{ij} \gamma_\mu Q_{j} \right] \left[ \overline{Q}_{i} (X_Q)_{ij} \gamma^\mu Q_{j} \right] \, , \end{eqnarray} where $\Lambda_{\rm NP}$ is some high energy scale and $z_1$ is the Wilson coefficient. $X_Q$ is a traceless hermitian matrix, transforming as an adjoint of $SU(3)_Q$ (or $SU(2)_Q$ for two generations). The contribution to $\Delta c,s\,=\,2$ transitions due to $X_Q$ is given by the misalignment between it and $\mathcal{A}_{u,d}$, and it is easy to see that this is equal to \begin{eqnarray} \label{2g_fv} \left| z_1^{D,K} \right|= \left| { X_Q} \times {\mathcal{A}_{u,d}} \right|^2/ |{\mathcal{A}_{u,d}}|^2 = \left| { X_Q} \times {\mathcal{\hat A}_{u,d}} \right|^2 \,, \end{eqnarray} where ${\mathcal{\hat A}_{u,d}} \equiv \mathcal{A}_{u,d} /\big|\mathcal{A}_{u,d} \big| \,.$ This result is manifestly invariant under a change of basis. Next we move to CPV \begin{eqnarray} \label{imz1k_def} \mathrm{Im}\left(z_1^{K,D}\right)=2\left(X_Q \cdot \hat J\right)\left( X_Q \cdot \hat J_{u,d} \right)\,, \end{eqnarray} where $\hat J\equiv \mathcal{A}_d \times \mathcal{A}_u/\left|\mathcal{A}_d \times \mathcal{A}_u\right|$ and $ \hat J_{u,d} \equiv \mathcal{\hat A}_{u,d}\times \hat J/\left| \mathcal{\hat A}_{u,d}\times \hat J \right|$. The above spurions and observables are easily described geometrically, say in the $\mathcal{\hat A}_d-\hat J-\hat J_d$ space, as shown in Fig.~\ref{fig:2g}. \begin{figure}[htb] \centering \includegraphics[width=2.33In]{2g_full.png} \caption{Flavor violation in the Kaon system induced by $X_Q$. The overall contribution to $K^0-\overline{K^0}$ mixing is given by the solid blue line. The CPV contribution, $\mathrm{Im}(z_1^K)$, is twice the product of the two solid orange lines, which are the projections of $X_Q$ on the $\hat J$ and $\hat J_d$ axes. Note that the angle between $\mathcal{A}_d$ and $\mathcal{A}_u$ is twice the Cabibbo angle, $\theta_{\rm C}$.} \label{fig:2g} \end{figure} To derive the weakest bound, we express $X_Q$ in terms of its components \begin{eqnarray}\label{XQ} X_Q=X^{u,d} \mathcal{\hat A}_{u,d}+X^J \hat J+X^{J_{u,d}} {\hat J}_{u,d} \,, \end{eqnarray} where we have $X^u=\cos 2\theta_{\rm C} X^d-\sin 2\theta_{\rm C} X^{J_d}$, $X^{J_u}=-\sin 2\theta_{\rm C} X^d-\cos 2\theta_{\rm C} X^{J_d}$ and $X^J$ remains invariant. Plugging the expression for $X_Q$ from Eq.~\eqref{XQ} into Eqs.~\eqref{2g_fv} and~\eqref{imz1k_def}, one easily reproduces the results of~\cite{Blum:2009sk} derived in a specific basis. A new condition for CPV is implied, exclusively related to $\Delta c,s\,=\,2$ processes and not to $\Delta c,s\,=\,1$ ones: \begin{eqnarray} \label{cpv_cond} X^{J_{u,d}} \propto \mathrm{tr} \left(X_Q [\mathcal{A}_{u,d},[\mathcal{A}_d,\mathcal{A}_u]] \right) \neq 0 \,, \end{eqnarray} while $X^J\neq0$ provides a necessary condition for all types of two generations CPV~\cite{Blum:2009sk}. The conditions are physically transparent and involve only observables, where the weakest bound on NP is derived for the ratio $X^d/X^{J_d}$ given a fixed amount of CPV, $X^J$. Note, however, that this new condition in Eq.~\eqref{cpv_cond} is only applicable to either the down or the up sector, while $X^J\neq0$ is universal. \section{Three Generations} For three generations, a simple 3D geometric interpretation does not naturally emerge anymore, as the relevant space is characterized by the eight Gell-Mann matrices. A useful approximation appropriate for third generation flavor violation is to neglect the masses of the first two generation quarks, where the breaking of the flavor symmetry is characterized by $[U(3)/U(2)]^2$~\cite{Kagan:2009bn}. It is especially suitable for the LHC, where it would be difficult to distinguish between light quark jets of different flavor. In this limit the CKM matrix is reduced to a real matrix with a single rotation angle between an active light flavor (say, the 2nd one) and the 3rd generation, \begin{eqnarray} \theta\cong \sqrt{\theta_{13}^2+\theta_{23}^2}\,, \end{eqnarray} where $\theta_{13}$ and $\theta_{23}$ are the corresponding CKM mixing angles. The other generation (the first one) decouples, and is protected by a residual $U(1)_Q$ symmetry~\cite{ourlong}. As before, we wish to analyze the flavor violation induced by $X_Q$ in a covariant form. The new contributions to $\Delta t,b\,=\,1$ transitions are characterized by \begin{eqnarray} \label{inclusive_decay} {\rm BR}(Q_3\to Q_{l})\propto \frac{4}{3} {\left| X_Q \times \mathcal{\hat A}_{u,d} \right|^2}\, , \end{eqnarray} where $Q_l$ stands for light doublets. We stick to the same definitions as in the two generation part, Eq.~\eqref{definitions}. In the $[U(3)/U(2)]^2$ limit we can covariantly identify four independent directions out of the eight generators space: $\mathcal{\hat A}_d$, $\hat J$, $\hat J_d$ and an additional one, $\hat J_Q \equiv -2\mathcal{\hat A}_d+\sqrt3\,\hat J\times \hat J_d$. Since $\mathcal{\hat A}_d$ and $\hat J_Q$ are not orthogonal, we replace the former with $\mathcal{\hat A}'_d\equiv \hat J\times \hat J_d$ (and a similar expression for the up sector). Note that $\hat J_Q$ corresponds to the conserved $U(1)_Q$ generator, so it commutes with both $\mathcal{A}_d$ and $\mathcal{A}_u$, and takes the same form when interchanging $d\leftrightarrow u$ ($\hat J$ also remains the same in both up and down bases, as in the two generations case). There are four additional directions, collectively denoted as $\hat {\vec D}$, which transform as a doublet of the CKM (2-3) rotation, and do not mix with the other directions. \section{Application~-- Third Generation Decays} We next use measurements of down type FCNC and LHC projection for top FCNC to derive a model independent bound on the corresponding NP scale. We focus on the following operator \begin{eqnarray} \label{3g_operator} O^h_{LL}= i \left[ \overline{Q}_i \gamma^\mu (X_Q^{\Delta F=1})_{ij} Q_j \right] \left[ H^\dagger \overleftrightarrow{D}_\mu H \right] + \mathrm{h.c.} \, , \end{eqnarray} which contributes at tree level to both top and bottom decays~\cite{Fox:2007in}\footnote{It is important to note that a given new physics model might generate different higher-dimensional operators via different types of processes. Therefore $X_Q$ is in general different for each operator, so we denote it specifically as $X_Q^{\Delta F=1}$ for the current case.}. We adopt the weakest limits on the coefficient of this operator, $C^h_{LL}$, derived in~\cite{Fox:2007in}: \begin{eqnarray} \label{exp_constraints} \begin{split} \mathrm{Br}&(B \to X_s\ell^+ \ell^-) \longrightarrow \left| C^h_{LL} \right|_b < 0.018 \left( \frac{\Lambda_{\rm NP}}{1\,\textrm{TeV}} \right)^2 \, , \\ \mathrm{Br}&(t\to (c,u)Z) \longrightarrow \left| C^h_{LL} \right|_t < 0.18 \left( \frac{\Lambda_{\rm NP}}{1\,\textrm{TeV}} \right)^2 \, , \end{split} \end{eqnarray} where the latter is based on the prospect for the LHC bound in the absence of signal, with 100\,fb$^{-1}$, and we define $r_{tb} \equiv \left| C^h_{LL} \right|_t/\left| C^h_{LL} \right|_b\,$. The NP contribution can be decomposed in the covariant bases \begin{eqnarray} \label{xq_param} \hspace*{-.25cm} X_Q \! = \! X'^{u,d} \mathcal{\hat A}'_{u,d}\hspace*{-.045cm} +\hspace*{-.045cm} X^J \hat J\hspace*{-.045cm} +\hspace*{-.045cm} X^{J_{u,d}} {\hat J}_{u,d}\hspace*{-.045cm} +\hspace*{-.045cm} X^{J_Q} \hat J_Q\hspace*{-.045cm} +\hspace*{-.045cm} X^{\vec D} \mathcal{\hat {\vec D}} . \end{eqnarray} The weakest bound is obtained, for a fixed length $L\equiv \left|X_Q\right|$, by finding a direction of $X_Q$ that minimizes the contributions to $\left| C^h_{LL} \right|_t$ and $\left| C^h_{LL} \right|_b$\,. It is clear, however, that directions that contribute to first two generations flavor and CPV at ${\cal O}\left(\lambda_{\rm C}\right)$ ($\lambda_{\rm C}\sim 0.23$) are strongly constrained. Thus, the resulting bounds would not correspond to the best alignment case. For example, when only $X^{J_Q}\neq0$, no third generation flavor violation is induced. However, switching back on the light quark masses, $X^{J_Q}$ (more precisely, a combination of $X^{J_Q}$ and $X'^d$) does induce flavor violation between the first two generations. At best it can be aligned with the down mass basis, so that it contributes to $\Delta c=1$ transition at ${\cal O}\left(\lambda_{\rm C}\right)$. The corresponding bound is~\cite{ourlong} \begin{eqnarray} \label{3g2g_constraint} L<0.59 \left( \frac{\Lambda_{\rm NP}}{1\,\textrm{TeV}} \right)^2; \ \ \Lambda_{\rm NP}>1.7 \, \mathrm{TeV} \, , \end{eqnarray} where the latter is for $L=1$. Similarly, it can be shown that $X^{\vec D}$ yields $2\to1$ transitions when the contributions to third generation decays are minimized. These cases, therefore, do not represent the best alignment scenario. The induced flavor violation is then given by \begin{eqnarray} \label{explicit} \frac{4}{3} \left| X_Q \times \mathcal{\hat A}_{u,d} \right|^2 = \left(X^J\right)^2+ \left(X^{J_{u,d}}\right)^2 \,, \end{eqnarray} and \begin{eqnarray} \label{explicit1} X^{J_{u}}=\cos 2\theta \,X^{J_{d}}+ \sin 2\theta \,X'^{d}\,. \end{eqnarray} From the above relations it is clear that $X^J$ contributes the same to both rates, so it should be set to zero for optimal alignment. Thus the best alignment is obtained by varying $\alpha$, defined by \begin{eqnarray} \tan\alpha \equiv X^{J_{d}}/X^d\,, \end{eqnarray} where $X^d$ is the coefficient of $\mathcal{\hat A}_d$, which is the generator that does not produce flavor violation among the first two generations to leading order (up to $\mathcal{O}(\lambda_{\rm C}^5)$). We now consider two possibilities: (i) complete alignment with the down sector; (ii) the best alignment satisfying the bounds of Eq.~\eqref{exp_constraints}, which gives the weakest unavoidable limit. The bounds for these cases are \begin{eqnarray} \label{3g_bounds} \mathrm{i)}& \hspace*{-.35cm}\alpha=0 \, , \ L<2.5 \left( \frac{\Lambda_{\rm NP}}{1\,\textrm{TeV}} \right)^2; \ \Lambda_{\rm NP}\!>0.63 \,(7.9)\, \mathrm{TeV} \, , \\ \mathrm{ii)}& \alpha=\frac{\sqrt{3}\,\theta}{1+r_{tb}} \, , \ L<2.8 \left( \frac{\Lambda_{\rm NP}}{1\,\textrm{TeV}} \right)^2; \ \Lambda_{\rm NP}\!>0.6 \, (7.6)\, \mathrm{TeV},\nonumber \end{eqnarray} as shown in Fig.~\ref{fig:3g_bounds}, where in parentheses we give the strong coupling bound, in which the coefficient of the operators in Eqs.~\eqref{o1} and~\eqref{3g_operator} is assumed to be $16 \pi^2$. Note that these are weaker than the bound in Eq.~\eqref{3g2g_constraint}. \begin{figure}[hbt] \centering \includegraphics[width=0.45\textwidth]{3g_bounds.png} \caption{Upper bounds on $L$ as a function of $\alpha$, coming from the measurements of flavor violating decays of the bottom and the top quarks, assuming $\Lambda_{\rm NP}=1$\,TeV.} \label{fig:3g_bounds} \end{figure} It is important to mention that the optimized form of $X_Q$ generates also $c \to u$ decay at higher order in $\lambda_{\rm C}$, which might yield stronger constraints than the top decay. In~\cite{ourlong} it is shown that the bound from the former is actually much {\emph{weaker}} than the one from the top, as a result of a $\lambda_{\rm C}^5$ suppression. Therefore, the LHC is indeed projected to \emph{strengthen} the model independent constraints. \section{Third Generation $\Delta F=2$ Transitions} Next we analyze $\Delta F=2$ processes, where for simplicity, we only consider complete alignment with the down sector, \begin{eqnarray} X_Q^{\Delta F=2}=L \mathcal{\hat A}_d \,, \label{xq2} \end{eqnarray} as the constraints from this sector are much stronger. This generates in the up sector $D^0-\overline{D^0}$ mixing and top flavor violation. Yet, there is no top meson, so we analyze instead the process $uu \to tt$, which is most appropriate for the LHC (and related to mixing by crossing symmetry). This process was studied in the literature in the context of different models (see \textit{e.g.}~\cite{Larios:2003jq,Kraml:2005kb,Gao:2008vv} and refs.~therein). It is observed through the dilepton mode, in which two same-sign leptons are produced from the top quarks. We emphasize that in this case the parton distribution functions of the proton strongly break the approximate $U(2)$ symmetry of the first two generations. Thus, a useful bound is obtained only from the operator involving up (and not charm) quarks. In order to estimate the prospect for the LHC bound on same-sign tops production, we calculated the $uu \to tt$ cross section using MadGraph/MadEvent~\cite{Alwall:2007st}, as a $t$ (or $u$) channel process mediated by a heavy vector boson, matched onto the operator in Eq.~\eqref{o1}. We then used the fact that the cross section times the integrated luminosity must be lower than 3 for a 95\% exclusion, in the absence of signal~\cite{PDG}. Adding an assumption of 1\% signal efficiency~\cite{Larios:2003jq}, after background reduction, we have \begin{eqnarray} z_1^{tt}<7.1 \times 10^{-3} \left( \frac{\Lambda_{\rm NP}}{1\,\textrm{TeV}} \right) ^2 \, , \end{eqnarray} for 100 fb$^{-1}$ at a center of mass energy of 14 TeV. The experimental constraint from CPV in the D system is~\cite{Gedalia:2009kh} \begin{eqnarray} \mathrm{Im}(z_1^D)<1.1 \times 10^{-7} \left( \frac{\Lambda_{\rm NP}}{1\,\textrm{TeV}} \right)^2 \, . \end{eqnarray} The contribution of $X_Q^{\Delta F=2}$ to these processes is calculated by applying a simple CKM rotation, and then taking $\mathrm{Im}\left[ \left(X_Q^{\Delta F=2} \right)^2_{12} \right]$ for CPV in $D$ mixing and $\left| \left( X_Q^{\Delta F=2} \right)_{13} \right|^2$ for $uu \to tt$. The resulting bounds are \begin{eqnarray} \begin{split} L&<1.8 \left( \frac{\Lambda_{\rm NP}}{1\,\textrm{TeV}} \right)\,; \ \ \Lambda_{\rm NP}>0.57 \, (7.2)\, \mathrm{TeV} \, , \\ L&<12 \left( \frac{\Lambda_{\rm NP}}{1\,\textrm{TeV}} \right)\,; \ \ \Lambda_{\rm NP}>0.08 \, (1.0)\, \mathrm{TeV} \, , \end{split} \end{eqnarray} for $D$ mixing and $uu \to tt$, respectively. Note that the latter bound depends on the quartic root of the cross section that was evaluated above, thus it is only mildly sensitive to that calculation and to the efficiency assumption. Anyway, in this case the existing bound is stronger than the one which will be obtained at the LHC for top quarks, as opposed to $\Delta F=1$ case considered above. \section{Supersymmetry} We now consider the application of our formalism to $\Delta F=2$ transitions in supersymmetry (constraints from $\Delta F=1$ processes are more involved, due to a richer operator structure, and discussed in~\cite{ourlong}). We use the approximation of quasi degenerate squark masses (see {\it e.g.}~\cite{Ciuchini:1998ix}), and consider the leading order in the expansion $\Delta\tilde m^2_{Q_2Q_1}$, ($\Delta\tilde m^2_{Q_iQ_j}$ is the mass-squared difference between the $i$th and $j$th squarks), where the level of degeneracy is much stronger~\cite{Blum:2009sk}. We arrive at the following expression for the length of $X_Q$ \begin{eqnarray} L=\frac{\alpha_{s}}{18}\,\sqrt{\frac{g(x)}{2}} \, \frac{\Delta\tilde m^2_{Q_3Q_1}}{\tilde m_Q^2}\,, \end{eqnarray} where $x=m_{\tilde{g}}^{2}/\tilde m_{Q}^{2}$, $m_{\tilde{g}}$ is the gluino mass and $g(x)$ is a known kinematic function~\cite{Ciuchini:1998ix}. Taking for concreteness, $\tilde m_Q=\left(2 m_{\tilde Q_1}+m_{\tilde Q_3}\right)/3$ (appropriate for models with only weak degeneracy~\cite{Raz:2002zx}), $\tilde m_Q = 100\mbox{\,GeV}$ and $m_{\tilde{g}}\approx \tilde m_{{Q}}$, which implies $g(1)=1$, we find \begin{eqnarray} \frac{\left|m_{\tilde Q_1}^2- m_{\tilde Q_3}^2\right|}{\left(2 m_{\tilde Q_1}+m_{\tilde Q_3}\right)^2} <0.45\left(\frac{\tilde m_{Q}}{100\mbox{\,GeV}}\right)^2\,. \end{eqnarray} \section{Warped Extra Dimension} Another example for a concrete model that is constrained by measurements is the Randall-Sundrum (RS) framework~\cite{Randall:1999ee}. When the fermions are allowed to propagate in the bulk, their localization yields mass hierarchies and mixing angles, thus addressing the flavor puzzle. The $\Delta F=2$ process is induced at tree level by a Kaluza-Klein (KK) gluon exchange. The $\Delta F=1$ operator in Eq.~\eqref{3g_operator} is generated, among others, via mixing between the SM Z and its KK excitations, which results in a non-diagonal coupling in the mass basis~\cite{Agashe:2004cp, Agashe:2006wa}. For simplicity, we only focus below on these contributions, as the others are of the same order~\cite{Agashe:2006wa}. For the $\Delta F=2$ case we have \begin{eqnarray} m_{\rm KK}=\Lambda_{\rm NP} \, , \quad X_Q\cong \frac{g_{s*}}{\sqrt{6}} \,\mathrm{diag} (f^2_{Q^1},f^2_{Q^2},f^2_{Q^3}) \, , \end{eqnarray} before removing the trace, where $g_{s*}$ is the dimensionless 5D coupling of the gluon ($g_{s*} \approx 3$ at one loop~\cite{Agashe:2008uz}) and the $f_{Q^i}$'s are the values of the quark doublets on the IR brane. These are related to each other through the CKM elements~-- $f_{Q^1,Q^2}/f_{Q^3} \sim V_{ub},V_{cb}$. The resulting limit is \begin{eqnarray} m_{\rm KK}>0.4 f^2_{Q^3} \; \mathrm{TeV} \, , \end{eqnarray} where $f_{Q^3}$ is typically in the range of 0.4-$\sqrt{2}$. For the $\Delta F=1$ process we find \begin{eqnarray} X_Q \cong g_{Z*}\, \delta g_Z\, \mathrm{diag}(f^2_{Q^1},f^2_{Q^2},f^2_{Q^3}) \, , \end{eqnarray} where $g_{Z*}$ is the dimensionless 5D coupling of the Z to left-handed up type quarks ($g_{Z*} \cong 1.2$ at one loop) and $\delta g_Z \cong \log (M_{\mathrm{Pl}}/\mathrm{TeV}) \left(m_Z/m_{\rm KK} \right)^2$ describes the non-universal coupling coming from mixing between the different Z states. The bound that stems from this is \begin{eqnarray} m_{\rm KK}>0.33 f^2_{Q^3} \; \mathrm{TeV} \, . \end{eqnarray} \section{Conclusions} We find that projected LHC bounds on $\Delta t=1$ processes enable us to provide a new model independent constraint on the strength of left-handed quarks flavor violation, even in the presence of general flavor alignment mechanisms. The projected bound on $\Delta t=2$ transitions from same sign tops production at the LHC is also studied. In this case a surprising result is that a stronger robust bound already exists due to the experimental constraint on CP violation in $D-\overline D$ mixing. We use our analysis to obtain new limits on supersymmetric and warped extra dimension models of alignment, which are rather weak (as a result of the weaker experimental constraints, compared to the first two generations~-- see {\it e.g.}\ in~\cite{Blum:2009sk}), but replacing practically non-existing current bounds. {\bf Acknowledgments } G.P. is supported by the Israel Science Foundation (grant \#1087/09), EU-FP7 Marie Curie, IRG fellowship and the Peter \& Patricia Gruber Award.
\section{Introduction}\label{sec:intro} Non-perturbative properties of non-abelian gauge theories, in particular color confinement, chiral symmetry breaking, and the existence of the mass gap, must be encoded in the structure of the ground state (vacuum) of these theories. It is natural, then, to look for evidence of these properties in the vacuum wavefunctional $\Psi_0[A]$ of quantized gauge field theory in some physical gauge. The simplest non-trivial setting is SU(2) gauge theory with no dynamical matter fields, and in one lower space dimension where the theory is super-renormalizable. Recently, we have proposed a very simple approximate form for the Yang--Mills vacuum wavefunctional in temporal gauge and $D=2+1$ dimensions \cite{Greensite:2007ij}:\footnote{Expressions below are assumed to be properly defined on a lattice, with lattice spacing serving as regulator, but for simplicity we will often use continuum notations.} \begin{eqnarray} \nonumber \Psi_0[A]&{=}&\label{eq:ourWF} \exp\Biggl[-{\textstyle{\frac{1}{2}}}\displaystyle\int d^2xd^2y\\ &\times&B^a(x){\displaystyle \left(\frac{1}{\sqrt{-{\cal D}^2- \lambda_0+m^2}}\right)_{xy}^{ab}}B^b(y)\Biggr]. \end{eqnarray} Here ${\cal D}^2$ is the covariant Laplacian in the adjoint representation, whose lowest eigenvalue is $\lambda_0$, $m$ is a constant with dimensions of mass proportional to $g^2\sim 1/\beta$, and ${B^a(x)}=F_{12}^a(x)$ is the color magnetic field strength. On the lattice, $-{\cal D}^2$ is given by \begin{eqnarray} \lefteqn{ \left({-{\cal D}^2}\right)^{ab}_{xy} =} \\ & & \displaystyle\sum_{k=1}^2 \left[ 2\delta^{ab}\delta_{xy} - {\cal U}^{ab}_k(x)\delta_{y,x+\hat{k}} -{\cal U}^{\dagger ba}_k(x-\hat{k})\delta_{y,x-\hat{k}}\right], \nonumber \end{eqnarray} where the ${\cal U}_k(x)$ are the link fields in the adjoint representation. The wavefunctional in Eq.~(\ref{eq:ourWF}) is reminiscent of Samuel's~\cite{Samuel:1996bt}, the difference being that in his proposal a single free parameter $m^2_0$ replaces our $(-\lambda_0+m^2)$ in the denominator. The reason for subtracting the lowest eigenvalue from the operator $(-{\cal D}^2)$ is that our numerical simulations indicate that the spectrum of this operator may be divergent in the continuum limit \cite{Greensite:2007ij}.\footnote{It is a little difficult to compare our wavefunctional directly with that of Karabali and Nair \cite{Karabali:1998yq,Karabali:2009rg}, because their proposed wavefunctional in new variables, when converted back to temporal gauge and the usual variables of gauge theory, is not gauge invariant. One must therefore suggest some gauge invariant extension. The simplest such extension, proposed in \cite{Karabali:1998yq} and investigated numerically in \cite{Greensite:2007ij}, does not seem to give the correct string tension. Other extensions are, however, possible.} The proposed vacuum wavefunctional (\ref{eq:ourWF}) has quite a few attractive properties: \begin{enumerate} \item In the free-field limit ($g\to 0$), the covariant Laplacian turns into an ordinary Laplacian, $\lambda_0$ and $m$ vanish, and $\Psi_0[A]$ becomes the well-known vacuum wavefunctional of electrodynamics: \begin{eqnarray} \nonumber \Psi_0[A]&=& \exp\Bigg\lbrace-{\textstyle{\frac{1}{2}}}\displaystyle\int d^2xd^2y\; [\partial_1 A_2^a(x)-\partial_2 A_1^a(x)]\\ &\times&\left(\frac{\delta^{ab}}{\sqrt{-\nabla^2}}\right)_{xy}[\partial_1 A_2^b(y)-\partial_2 A_1^b(y)]\Bigg\rbrace. \end{eqnarray} \item The expression~(\ref{eq:ourWF}) is a good approximation to the true vacuum also in a completely different corner of the configuration space, namely if we restrict to fields constant in space and varying only in time. In the large-volume limit the solution of the Yang--Mills Schr\"odinger equation in that case is, up to $1/V$ corrections: \begin{equation} \Psi_0 =\exp\left[\displaystyle -{\textstyle{\frac{1}{2}}}gV\frac{(\vec{A_1}\times\vec{A_2})\cdot(\vec{A_1}\times\vec{A_2})} {\sqrt{\vec{A}_1\cdot\vec{A}_1+\vec{A}_2\cdot\vec{A}_2}} \right], \end{equation} and exactly the same expression follows from (\ref{eq:ourWF}) assuming $\vert g \vec{A}_{1,2}\vert \gg m, \sqrt{\lambda_0}$. \item If we divide the field strength $B^a(x)$ into ``fast'' and ``slow'' components, the part of the (squared) vacuum wavefunctional that depends only on $B_\mathrm{slow}$ is \begin{equation}\label{eq:dim_red} \vert\Psi_0\vert^2\approx\displaystyle\exp\left[ -\frac{1}{m}\int d^2x\; B^a_\mathrm{slow}(x)\;B^a_\mathrm{slow}(x)\right]. \end{equation} Such a form is expected on the basis of dimensional-reduction arguments \cite{Greensite:1979yn,Olesen:1981zp,Ambjorn:1984mb}: it is exactly the probability measure for Yang--Mills theory in two Euclidean dimensions, which (i)~is confining for $m>0$, and (ii) exhibits Casimir scaling for string tensions of all color-charge representations. The fundamental string tension is easily computed as $\sigma_F(\beta)=3 m/(4\beta)$. The last expression can be used to fix the value of $m$ in Eq.~(\ref{eq:ourWF}) at a given $\beta$ from the known value of $\sigma_F(\beta)$. \item Confinement requires $m$ to be larger than zero. If one takes $m$ in the wavefunctional (\ref{eq:ourWF}) as a variational parameter and computes (approximately) the expectation value of the Yang--Mills Hamiltonian, one finds that a non-zero (and finite) value of $m$ is energetically preferred. \item If we fix the mass $m$ in the wavefunctional to get the right string tension $\sigma_F$ at a given $\beta$, we can test our proposal by calculating \textit{e.g.}\ the mass gap of the theory. We have proposed a recursive procedure for generating independent lattice configurations with the probability distribution given by the square of the wavefunctional (\ref{eq:ourWF}) (see Ref.\ \cite{Greensite:2007ij} and Sec.\ \ref{sec:recursion} for details). We call two-dimensional lattice configurations obtained in this way ``recursion lattices''. One can compute observables with these lattices, and compare the results with corresponding values obtained from ``Monte Carlo lattices'', \textit{i.e.}\ two-dimensional slices of lattices generated in a full $D=3$ lattice Monte Carlo simulation. It turns out that, given the asymptotic string tension as input, the mass gap comes out fairly accurately from our wavefunctional. The discrepancies with the Monte Carlo results of Meyer and Teper \cite{Meyer:2003wx} for the $0^+$ glueball mass are at the level of at most a few ($<6$) percent. \item The dimensional reduction form (\ref{eq:dim_red}) at large distances implies an area law fall-off for large Wilson loops, and also Casimir scaling of higher-representation Wilson loops. The question is then how Casimir scaling turns into $N$-ality dependence at large distances, \textit{i.e.}\ how color screening enters in this setting. There are indications that terms needed for color screening might be contained in our simple wavefunctional and would appear as corrections to the dimensional-reduction form~(\ref{eq:dim_red}). \end{enumerate} In this article we proceed further in comparing quantities derived from our proposed vacuum wavefunctional, with those obtained from full Monte Carlo simulations of the Yang--Mills theory, by computing the Coulomb-gauge ghost propagator and the color-Coulomb potential. The translation between temporal gauge and the minimal Coulomb gauge was discussed in detail in Ref.\ \cite{Greensite:2004ke}, but the conclusion (\textit{cf.\/}~Sec.~\ref{sec:translation}) is simply that the Yang--Mills wavefunctional in Coulomb gauge can be obtained by restricting the temporal-gauge wavefunctional to transverse gauge fields. In Sec.~\ref{sec:recursion} we review the recursion procedure used to generate lattice configurations with probability weighting given by the square of the wavefunctional~(\ref{eq:ourWF}). Our numerical results are presented in Sec.~\ref{sec:results}, first for the Coulomb-gauge ghost propagator in Sec.~\ref{subsec:ghost}, and then for the color-Coulomb potential in Sec.~\ref{subsec:potential}. The latter section also contains a discussion of subtleties in the determination of the color-Coulomb potential, due to exceptional lattice configurations. Our conclusions are briefly summarized in Sec.~\ref{sec:conclusions}. \section{From temporal to Coulomb gauge}\label{sec:translation} The proposed wavefunctional, Eq.~(\ref{eq:ourWF}), is formulated as an approximate solution of the Yang--Mills Schr\"odinger equation in the temporal gauge. To compute quantities in Coulomb gauge we need to find its Coulomb-gauge equivalent. This is an easy task, due to the fact that both the temporal and the Coulomb gauge are compatible with a Hamiltonian formulation and a physical transfer matrix. In temporal gauge, $A_0=0$, in $D=2+1$ dimensions the continuum Hamiltonian has the simple canonical form \begin{equation}\label{eq:TGHamiltonian} H = {\textstyle{\frac{1}{2}}} \int d^2 x \left[ \sum_{i=1}^2 E_i^a(x)^2 + B^{a}(x)^2 \right], \end{equation} where the color-electric fields $E_i^a(x)=-i[\delta/\delta A^a_i(x)]$ are operators canonically conjugate to the spatial components of the vector potential. The temporal gauge still possesses invariance under space-dependent, time-independent gauge transformations $\Omega(x)$. The generator of local infinitesimal space-dependent gauge transformations, in the absence of external color sources, is given by \begin{equation}\label{eq:generator} G^a(x)=-\sum_{i=1}^2 {\cal{D}}^{ab}_i[A]\;E^b_i(x), \end{equation} ${\cal{D}}_i[A]$ is the covariant derivative. Physical wavefunctionals are required to satisfy the Gauss law constraint \begin{equation}\label{eq:Gauss} G^a(x)\Psi[A]=\sum_{i=1}^2\left( \delta^{ac} \partial_i + g \epsilon^{abc} A^b_i \right) {\delta \over \delta A^c_i }\Psi[A] = 0, \end{equation} which means that the wavefunctional must be invariant under infinitesimal gauge transformations. Due to the local gauge invariance of wavefunctionals in the temporal gauge, the volume of the gauge group can be extracted from inner products of wavefunctionals \begin{equation}\label{eq:inner} \langle\Psi_1\vert\Psi_2\rangle=\int[DA]\;\Psi_1^\ast[A]\Psi_2[A] \end{equation} via the Faddeev--Popov procedure \cite{Zwanziger:2003cf}. We parametrize configurations $A$ by $A={}^{\Omega^{-1}} A_\perp$, where $A_\perp$ is the representative of $A$ in the minimal Coulomb gauge, \textit{i.e.}\ $A_\perp$ is transverse, $\nabla\cdot A_\perp=0$, and belongs to the fundamental modular region (FMR), $A_\perp\in \Lambda$; $\Omega$~denotes the gauge transformation that brings the configuration $A$ to the minimal Coulomb gauge. The Faddeev--Popov formula then gives \begin{equation}\label{eq:innerFP} \langle\Psi_1\vert\Psi_2\rangle=\int_\Lambda\;[DA_\perp]\;\det {\cal{M}}[A_\perp]\;\Psi_1^\ast[A_\perp]\Psi_2[A_\perp], \end{equation} where ${\cal{M}}[A_\perp]=-\nabla\cdot {\cal{D}}[A_\perp]$ is the Faddeev--Popov operator, symmetric and positive-definite for $A_\perp\in\Lambda$. The right side of Eq.~(\ref{eq:innerFP}) is the proper expression for the inner product in the minimal Coulomb gauge. This means that in the operator formalism, the minimal Coulomb gauge is a gauge fixing within the temporal gauge of the remnant local gauge invariance. The wavefunctional in Coulomb gauge is the restriction of the wavefunctional in temporal gauge to transverse fields in the FMR: \begin{equation}\label{eq:WFcoul} \Psi^{\mathrm{Coulomb}}[A_\perp]=\Psi[A_\perp],\qquad A_\perp\in\Lambda. \end{equation} (The conditions $A_0=0$ and $\nabla\cdot A=0$ cannot both be imposed at all times, but can be imposed at one fixed time, and this is all that is required.) The vacuum expectation value of an operator $Q$ in Coulomb gauge can then be computed from \begin{equation} \langle Q\rangle =\langle\Psi_0^{\mathrm{Coulomb}}\vert Q[A_\perp]\vert\Psi_0^{\mathrm{Coulomb}}\rangle \end{equation} Inverting the Faddeev-Popov gauge fixing takes us to \begin{equation} \langle Q \rangle = \langle\Psi_0\vert Q\left[{}^\Omega \! A\right]\vert\Psi_0\rangle, \label{eq:vev} \end{equation} \textit{i.e.}\ we generate configurations following the probability distribution $\Psi_0^2$, transform them to the Coulomb gauge, and evaluate the observable $Q$ in the transformed configuration. From the path-integral representation of the vacuum state, we may also go from (\ref{eq:vev}) to \begin{equation} \langle Q \rangle = \left\langle Q\left[{}^{\Omega'}\!\! A({\bf x},t=0)\right] \right\rangle \end{equation} where the right hand side is the expectation value obtained in $D=3$ Euclidean dimensions, and $\Omega'$ is the gauge transformation which takes the gauge field on a $t=0$ time-slice into Coulomb gauge. \section{Generation of ``recursion lattices''}\label{sec:recursion} To assess the consequences of the proposed wavefunctional, Eq.\ (\ref{eq:ourWF}), one would like to generate lattice configurations with probability distribution given by the square of the wavefunctional\footnote{We have absorbed the coupling $g$ into the definition of $A_k$, which accounts for the factor of $1/g^2$ in the exponent in Eq.~(\ref{eq:g2}).} \begin{eqnarray} \lefteqn{P[A] = \left\vert\Psi_0[A]\right\vert^2} \nonumber \\ &=& \exp\left[- {1\over g^2}\int d^2x d^2y ~ B^a(x) {\cal{K}}^{ab}_{xy}[A] B^b(y) \right], \label{eq:g2} \end{eqnarray} where \begin{equation}\label{eq:kernel} {\cal{K}}_{xy}^{ab}[A] = \left({1 \over \sqrt{-{\cal{D}}^2 - \lambda_0 + m^2}} \right)^{ab}_{xy} \end{equation} We have proposed, in Ref.\ \cite{Greensite:2007ij}, an iterative procedure to achieve this goal. Here we summarize, for completeness and reader's convenience, its most important points. We define a probability distribution of gauge fields $A$ with the kernel ${\cal{K}}$ depending on the background field $A'$: \begin{eqnarray} &&P\left[A;{\cal{K}}[A']\right] \nonumber\\ &\sim& \exp\left[-{1\over g^2}\int d^2x d^2y ~ B^a(x) {\cal{K}}^{ab}_{xy}[A'] B^b(y) \right]. \label{eq:PAK} \end{eqnarray} Assuming the variance of the kernel ${\cal{K}}$ is small among thermalized configurations gauge-fixed to an appropriate gauge, we can approximate $P[A]$ by \begin{eqnarray} P[A] &\approx& P\left[A,\langle {\cal{K}} \rangle \right] = P\left[A, \int DA' ~ {\cal{K}}[A'] P[A'] \right] \nonumber \\ &\approx& \int DA' ~ P\left[A,{\cal{K}}[A']\right] P[A'], \label{eq:PA} \end{eqnarray} and solve the equation iteratively \begin{eqnarray} P^{(1)}[A] &=& P\left[A;{\cal{K}}[0]\right], \nonumber \\ \nonumber\dots &&\\ P^{(n+1)}[A] &=& \int DA' ~ P\left[A;{\cal{K}}[A']\right] P^{(n)}[A']. \end{eqnarray} In practice, we work on a two-dimensional lattice in an axial gauge ($A_1=0$), which enables one to change variables from $A_1$ and $A_2$ to $B$ cheaply, without introducing a field-dependent jacobian. Initially, we set also $A_2=0$. Then the iterative procedure consists of the following steps: \begin{itemize} \item[(i)] Given $A_2$, set $A'_2=A_2$. \item[(ii)] The probability $P\left[A;K[A']\right]$ is gaussian in $B$, diagonalize $K[A']$ and generate new $B$-field stochastically. \item[(iii)] From $B$, calculate $A_2$ in axial gauge, and compute everything of interest. \item[(iv)] Go back to step (i), repeat as many times as necessary. \end{itemize} Lattice configurations generated by this procedure are referred to as ``recursion lattices.'' It turns out that the procedure converges rapidly, after $\cal{O}$(10) iterations (cycles above), and the assumption about a small variance of $\cal{K}$ among configurations is supported \textit{a posteriori} by the absence of large fluctuations of the spectrum of $\cal{K}$ evaluated on individual recursion lattices. \section{Coulomb-gauge results}\label{sec:results} Confinement exists, of course, in any gauge, but in some gauges the phenomenon may be easier to understand than in others. Coulomb gauge has received some attention, following the seminal works of Gribov \cite{Gribov:1977wm} and Zwanziger \cite{Zwanziger:1998ez}, who argued that the low-lying spectrum of the Faddeev--Popov operator in Coulomb gauge probes properties of non-abelian gauge fields that are crucial for the confinement mechanism. The ghost propagator in Coulomb gauge and the color-Coulomb potential are directly related to the inverse of the Faddeev--Popov operator, and play a role in various confinement scenarios. In particular, the color-Coulomb potential represents an upper bound on the physical potential between a static quark and antiquark, which means that a confining color-Coulomb potential is a necessary condition to have a confining static quark potential \cite{Zwanziger:2002sh}. It is therefore an important check on the validity of our proposed vacuum wavefunctional, to see how well it can reproduce the values of Coulomb-gauge observables, such as the ghost propagator and color-Coulomb potential, that can be obtained by standard lattice Monte Carlo techniques. \subsection{Ghost propagator}\label{subsec:ghost} The ghost propagator in Coulomb gauge is given by the inverse of the Faddeev--Popov operator: \begin{eqnarray} \nonumber G(R)&=&\left.\left\langle\left({\cal{M}}[A]^{-1}\right)^{aa}_{xy}\right\rangle\right\vert_{\vert x-y\vert=R}\\ &=&\left.\left\langle\left(-\frac{1}{\nabla\cdot{\cal{D}}[A]}\right)^{aa}_{xy}\right\rangle\right\vert_{\vert x-y\vert=R}. \label{eq:ghost_prop} \end{eqnarray} In $D=2+1$ dimensions, with SU(2) gauge group and lattice link matrices parametrized by \begin{equation} U_\mu(x)=b_\mu(x)\mathbf{1}+i \sigma^c a^c_\mu(x),\quad b_\mu(x)^2+a^c_\mu(x)^2=0, \end{equation} the lattice Faddeev--Popov operator is given by (assuming the lattice version of the Coulomb-gauge condition $\nabla\cdot A=0$ is satisfied) \begin{eqnarray} \nonumber {\cal{M}}^{ab}_{xy}&=& \delta^{ab} \sum_{k=1}^2\left\{ \delta_{xy} \left[b_k(x) + b_k(x-\hat{k})\right]\right.\\ &-& \left. \delta_{x,y-\hat{k}} b_k(x) - \delta_{y,x-\hat{k}} b_k(y) \right\}\label{eq:FP}\\ \nonumber &-& \epsilon^{abc} \sum_{k=1}^3\left\{ \delta_{x,y-\hat{k}} a^c_k(x) - \delta_{y,x-\hat{k}} a^c_k(y) \right\}. \end{eqnarray} This operator is symmetric and positive-definite anywhere in the Coulomb-gauge Gribov region, \textit{i.e.}\ for configurations that are local minima of the quantity \begin{equation} {\cal{R}}[U]=-\sum_x\sum_{k=1}^2 \mbox{Tr}[U_k(x)]. \label{eq:Rcoul} \end{equation} More precisely, on a lattice with periodic boundary conditions the Faddeev--Popov operator possesses three trivial zero eigenvalues, the eigenmodes being independent of $x$. The existence of these modes is a consequence of the fact that, even apart from Gribov copies, the Coulomb-gauge condition does not fix the gauge completely. There is a remnant global symmetry such that if a set of link matrices $U_k(x)$ satisfies the condition, so does $U_k'(x)=\Omega\;U_k(x)\;\Omega^\dagger$, where $\Omega$ is a space-independent SU(2) group element. The inversion of the Faddeev--Popov operator, Eq.~(\ref{eq:FP}), needed to compute the ghost propagator (\ref{eq:ghost_prop}), is therefore performed in the subspace orthogonal to the trivial zero eigenmodes. At lattice coupling $\beta=6$ on a $24^2$ lattice, and $\beta=9$ on a $32^2$ lattice, we have computed the Coulomb-gauge ghost propagator separately in two ensembles of lattice configurations generated in two different ways: \begin{itemize} \item[(i)] \textit{recursion lattices} are generated by the procedure described in Section \ref{sec:recursion}; the mass parameter $m$ in the wavefunctional (\ref{eq:ourWF}), at each lattice coupling, was chosen to reproduce the value of the string tension at the same coupling as given by Ref.\ \cite{Meyer:2003wx}; \item[(ii)] \textit{Monte Carlo lattices} are generated by Monte Carlo simulations of euclidean SU(2) lattice gauge theory in $D=3$ dimensions with the standard Wilson action. From each thermalized configuration only a single (random) space slice at fixed euclidean time was chosen. The probability measure for such lattice time-slices, when transformed to Coulomb gauge, is given by the true Coulomb-gauge vacuum wavefunctional (i.e.\ the vacuum state of the corresponding lattice transfer matrix). \end{itemize} There were 1000 lattice configurations in each ensemble, at each of the couplings $\beta=6$ and $9$. Each two-dimensional lattice configuration was fixed to the Coulomb gauge by minimizing the quantity $\cal{R}$, Eq.~(\ref{eq:Rcoul}), via the usual (over)relaxation method.\footnote{It should be noted that this procedure returns a gauge copy in the Gribov region; no procedure for fixing to the fundamental modular region exists. However, we feel that there is nothing sacred (or more physical) about Gribov copies in the fundamental modular region, and will be satisfied with local, rather than global minima of $\cal{R}$.} Then, the inverse of the Faddeev--Popov operator was computed using the standard linear-algebra tools (\texttt{Octave}). This enabled us to determine the ghost propagator directly in the coordinate representation, in contrast to most other lattice investigations which generally determine propagators in the momentum representation (see \textit{e.g.\/} Refs.\ \cite{Quandt:2008zj,Nakagawa:2009zf}). The results for our largest coupling ($\beta=9$) and lattice size ($32^2$) are displayed in Fig.~\ref{fig:propagator} (right). The agreement between recursion and Monte Carlo lattices is really quite striking, with the differences between the displayed data sets being comparable to the size of the symbols. The same agreement is observed also for $\beta=6$ on $24^2$ lattice, \textit{cf.\/} Fig.~\ref{fig:propagator} (left). \begin{figure*}[bth!] \centering \begin{tabular}{c c} \includegraphics[width=0.48\hsize]{propagator_b6_l24_all} & \includegraphics[width=0.48\hsize]{propagator_b9_l32_all} \end{tabular} \caption{The Coulomb-gauge ghost propagator: (left) $\beta=6$ on $24^2$ lattice, (right) $\beta=9$ on $32^2$ lattice.} \label{fig:propagator} \end{figure*} \subsection{Color-Coulomb potential}\label{subsec:potential} The potential between a static quark and antiquark located at points $x$ and $y$, respectively, is proportional to:\footnote{We have omitted below the normalization factor $g^2 C_F/d_A$, where $C_F$ is the eigenvalue of the quadratic Casimir operator in the fundamental color representation and $d_A$ is the dimension of the adjoint representation of the color group, and nevertheless call the above quantity the color-Coulomb potential. The normalization factor would be needed if we intended to get a numerically accurate value of the Coulomb string tension, which was not our goal in the present study.} \begin{eqnarray} V(R)&=&\left.-\left\langle\left({\cal{M}}[A]^{-1}(-\nabla^2){\cal{M}}[A]^{-1}\right)^{aa}_{xy}\right\rangle\right\vert_{\vert x-y\vert=R}\\ \nonumber &=&\left.-\left\langle\left(\frac{1}{\nabla\cdot{\cal{D}}[A]}(-\nabla^2)\frac{1}{\nabla\cdot{\cal{D}}[A]}\right)^{aa}_{xy}\right\rangle\right\vert_{\vert x-y\vert=R}. \label{eq:Coulomb_potential} \end{eqnarray} Provided we know the inverse of the Faddeev--Popov operator in a configuration (on the subspace orthogonal to trivial zero modes, see the preceding section), the computation of the potential is quite straightforward. The result is shown in Fig.~\ref{fig:potential}. After what we have seen in Section \ref{subsec:ghost}, the figures for potentials come as a surprise. In the case of the ghost propagator, there was almost no difference between recursion and Monte Carlo lattice ensembles. Now we observe quite strong differences. The origin of these differences can fortunately be identified. Both in the Monte Carlo and recursion ensemble there exist ``exceptional'' configurations in which the lowest nontrivial eigenvalue of the Faddeev--Popov operator $\mu_0$ is still positive, but extremely small. It means that in the space of gauge-equivalent configurations there exist a valley along which the minimized quantity $\cal{R}$, Eq.~(\ref{eq:Rcoul}), almost does not change. Consequently, these configurations were always rather difficult to gauge-fix to Coulomb gauge.\footnote{A similar type of ``exceptional'' hard-to-gauge-fix configurations was encountered \textit{e.g.\/} by the Berlin group\ \cite{Voigt:2008rr}.} The existence of these exceptional configurations does not have a crucial impact on the ghost propagator, since its definition contains a single power of the inverse Faddeev--Popov operator, but their influence is strongly amplified in color-Coulomb potentials, where the inverse appears twice. \begin{figure*}[tbh!] \centering \begin{tabular}{c c} \includegraphics[width=0.48\hsize]{potential_b6_l24_all} & \includegraphics[width=0.48\hsize]{potential_b9_l32_all} \end{tabular} \caption{The color-Coulomb potential computed from all measured configurations: (left) $\beta=6$ on $24^2$ lattice, (right) $\beta=9$ on $32^2$ lattice.} \label{fig:potential} \end{figure*} We will illustrate this point on the data for $\beta=9$ on $32^2$ lattice. One can evaluate the values of the potential in a single lattice configuration. The lowest eigenvalue of the Faddeev--Popov operator directly influences the absolute value of the potential at the origin, $\vert V(0)\vert$, so one can classify configurations by their values of $\vert V(0)\vert$, and evaluate average potentials from sets of configurations satisfying a number of cuts: $\{\vert V(0)\vert<\kappa_i, i=1,2,\dots, K\}$. Figure \ref{fig:v0_b9_l32b} (left) displays values of $\vert V(0)\vert$ in the individual configurations from the recursion and Monte Carlo ensembles. One can see that the majority of configurations lie in the range between about 2 to 20, but there are rare instances of configurations above 100 (among Monte Carlo lattices) or even 400 (among recursion lattices). The frequency of these ``exceptional'' configurations in the ensemble is very small, but such configurations give rise to tremendous fluctuations in the measured average values of the potential. The distribution of configurations according to $\vert V(0)\vert$ is further illustrated in Fig.~\ref{fig:v0_b9_l32b} (right). \begin{figure*}[bth!] \centering \begin{tabular}{c c} \includegraphics[width=0.48\hsize]{v0_b9_l32}& \includegraphics[width=0.48\hsize]{no_with_limits} \end{tabular} \caption{Left: $\vert V(0)\vert$ in the individual configurations from the recursion and Monte Carlo ensembles. Right: The fraction of configurations with $\vert V(0)\vert<\kappa$. The top curve is for Monte Carlo lattices, the bottom one for recursion lattices. The area between the curves is colored. } \label{fig:v0_b9_l32b} \end{figure*} The $\kappa$-dependence of the average value of the magnitude of the color-Coulomb potential $V(R)$ at $R=0$, evaluated in subsets of Monte Carlo and recursion lattices with $\vert V(0)\vert<\kappa$, is displayed in Fig.~\ref{fig:v0_vs_kappa}. At $\beta=9$ (right panel), the average values in both ensembles agree reasonably at least for $\kappa\le{\cal{O}}(10)$. Such a cut on $\vert V(0)\vert$ is satisfied by about 85\% Monte Carlo lattices, and almost 80\% recursion lattices (\textit{cf.\/} Fig.~\ref{fig:v0_b9_l32b}, right). The severity of the problem of ``exceptional'' configurations is seen even more dramatically in the data for $\beta=6$ on $24^2$ lattice (Fig.~\ref{fig:v0_vs_kappa}, left panel). Here the close agreement of the average $\vert V(0)\vert$ between Monte Carlo and recursion lattices persists for the cut-off $\kappa$ at least ${\cal{O}}(100)$ (satisfied by 99\% of lattices); then, a single recursion lattice with extremely high value of $\vert V(0)\vert$ (\textit{i.e.}\ extremely small lowest eigenvalue of the Coulomb-gauge Faddeev--Popov operator) completely distorts the picture, and causes the disagreement between color-Coulomb potentials and the huge errorbars seen in Fig.~\ref{fig:potential}. If the color-Coulomb potential is evaluated in Monte Carlo and recursion lattices with the same (not too high) cuts applied in both ensembles, the fluctuations due to rare configurations, observed in Fig.~\ref{fig:potential}, are tamed. In Fig.~\ref{fig:potential_b9_l32_cuts} we show the results for $\beta=6$ with $\kappa=100$, and for $\beta=9$ with $\kappa=10$. The potentials are linearly rising over a certain range of distances, and agree quite well in recursion and Monte Carlo lattices. The agreement deteriorates somewhat when the value of the cut $\kappa$ is increased, but we believe that approximate agreement would be restored with sufficient (but obviously huge) statistics if the sets of configurations in high $\vert V(0)\vert$ bins were equally populated. \begin{figure*}[bth!] \centering \begin{tabular}{c c} \includegraphics[width=0.48\hsize]{v0_vs_kappa_b6_l24}& \includegraphics[width=0.48\hsize]{v0_vs_kappa_b9_l32} \end{tabular} \caption{The average of $\vert V(0)\vert$ evaluated in subsets of configurations satisfying the condition $\vert V(0)\vert<\kappa$: (left) $\beta=6$ on $24^2$ lattice, (right) $\beta=9$ on $32^2$ lattice.} \label{fig:v0_vs_kappa} \end{figure*} \begin{figure*}[bth!] \centering \begin{tabular}{c c} \includegraphics[width=0.48\hsize]{potential_b6_l24_l100}& \includegraphics[width=0.48\hsize]{potential_b9_l32_l10} \end{tabular} \caption{The color-Coulomb potential computed from configurations with cuts: (left) $\beta=6$ on $24^2$ lattice, $\kappa=100$, (right) $\beta=9$ on $32^2$ lattice, $\kappa=10$.} \label{fig:potential_b9_l32_cuts} \end{figure*} \section{Conclusions}\label{sec:conclusions} The results presented in this paper strengthen our confidence that the vacuum wavefunctional for the temporal-gauge SU(2) Yang--Mills theory in $D=2+1$ dimensions, Eq.~(\ref{eq:ourWF}), is a fairly good approximation to the true ground state of the theory. The evidence supplied by Ref.\ \cite{Greensite:2007ij} and summarized in Section\ \ref{sec:intro} has been now augmented by measurement of two quantities which play a crucial role in understanding confinement in Coulomb gauge: \begin{enumerate} \item The ghost propagator in Coulomb gauge, computed from an ensemble of recursion lattices (derived from our approximate wavefunctional), and from an ensemble of Monte Carlo lattices, comes out to be practically identical in both ensembles. \item If the same cuts on ``exceptional'' configurations are applied in both ensembles, then the color-Coulomb potential from recursion lattices is also very close to that determined from Monte Carlo lattices. Both potentials are linearly rising over a range of distances (until the effects of lattice periodicity become important). However, one would need to considerably increase the statistics to ensure approximately equal population of exceptional configurations in both ensembles, to convincingly prove that the deviation between potentials in these two ensembles stays tiny even after the cuts are removed. \end{enumerate} While we do not claim that Eq.~(\ref{eq:ourWF}) is exact (it is surely only an approximation to the true ground state), it has now passed a number of very non-trivial tests. More tests are in progress, and will be reported on at a later time. The extension of our proposed vacuum wavefunctional to $D=3+1$ dimensions is straightforward; it simply involves replacing the product $B^a(x) B^b(y)$ in Eq.~(\ref{eq:ourWF}) by $F_{ij}^a(x) F_{ij}^b(y)$, and replacing two-dimensional integrations by three-dimensional integrals. It remains true in $3+1$ dimensions that the resulting wavefunctional is exact in the free-field limit, and approximately solves the Yang--Mills Schr\"odinger equation in the zero-mode, strong-field limit (see Ref.\ \cite{Greensite:2007ij} for details). However, going to three space dimensions brings up complications associated with the Bianchi constraint. Because of that constraint, numerical simulation of the wavefunctional along the lines discussed in Section \ref{sec:recursion} is much more challenging in $D=3+1$ than in $D=2+1$, and a different approach to testing our proposal is probably required.\footnote{For some recent work using the dimensionally-reduced form of the 3+1 dimensional vacuum wavefunctional, \textit{cf.\/} Ref.~\cite{Quandt:2010yq}.} \acknowledgments{This research is supported in part by the U.S.\ Department of Energy under Grant No.\ DE-FG03-92ER40711 (J.G.), the Slovak Grant Agency for Science, Project VEGA No.\ 2/0070/09, by ERDF OP R\&D, Project CE QUTE ITMS 26240120009, and via CE SAS QUTE (\v{S}.O.).}
\section{Introduction} Quantum information processing relies on the capacity of sustaining coherence as well as the entanglement among different parties. However, many systems proposed for implementing such protocols are naturally subjected to local and unavoidable dissipation such as ion traps \cite{review ion}, cavity QED systems \cite{QED}, and atomic ensembles~\cite{atomicE}. Needless to say, this usually works against their quantum efficiency through the mechanism of decoherence. Different strategies have been designed to partially protect or restore this coherence, such as entanglement distillation~\cite{Distill}, quantum repeaters~\cite{Repeat} or feedback~\cite{Feedback}, even though all these proposals present combinations of different costs like the increase of the amount of resources to create the necessary redundancy or the need to perform multi-qubit, \textit{i.e.} non-local, operations. In this paper we show a fully deterministic and \emph{local} scheme that counteracts the unavoidable action of dissipation. It relies on three basic elements: the capacity to monitor local environments detecting whether one or none excitation has leaked into each local reservoir; the possibility to rapidly feedback the excitation into the lossy qubit; and the use of three-level systems to encode qubits. Note that even though the used encoding implies some sort of redundancy, the scheme is much less demanding than the usual quantum error correction codes for dissipative systems~\cite{Feedback} because it adds only one local extra level to each part and does not require any extra entanglement or global operation. We discuss applications of this idea to improve the efficiency of quantum repeater protocols and to produce longer lasting coherence and/or entanglement for quantum information storage, teleportation or swapping. We also show that it can represent a physical implementation of the optimum singlet conversion protocol~\cite{Vidal}. Let us consider a global system composed of internal subsystems that are weakly coupled to their own local reservoirs. These couplings should respect typical Markov and Born approximations and one should be able to read information from each environment regarding the emission of single excitations from the respective subsystem in a time scale much shorter than the subsystem decay rate and still much larger than the correlation times of its respective reservoir. We then show that, within the framework of the well known quantum trajectories technique~\cite{RevTrajec}, the monitoring of the local environments plus the classical communication of the obtained results (hence, the first element of the scheme) is already enough to enhance the quantum communication efficiency. However, in the common case of qubits encoded in two-level systems, only the so-called no-jump trajectory, obtained when no environment is disturbed, is of use. In fact, the scheme starts to fail as soon as any single qubit decays, since its particular state is projected to the low energy level and factored out of the rest of the system. In order to solve this problem, we introduce the two remaining elements to show that encoding the qubits in three-level systems, combined with feedback on individual systems powers up the scheme to the limiting point of completely suppressing the action of the local environments only through local operations. \section{Monitoring the environment} To start with, let us take a look at how environment-monitoring can enhance a simple quantum teleportation scheme for which Alice and Bob need to share a pair of maximally entangled qubits. Consider a initial Bell state given by $|\Psi_0\rangle = \frac{|10\rangle + |01\rangle}{\sqrt{2}}$ and that each qubit is under the action of a local spontaneous emission reservoir. Without external monitoring, the joint state will become a mixture and the pre-existing entanglement will eventually vanish. The time evolution of the whole system is described by the following master equation: \begin{eqnarray} \frac{d \rho}{dt} = -\frac{\gamma}{2} (\hat{a}^\dagger \hat{a} \rho + \rho \hat{a}^\dagger \hat{a}) + \gamma \hat{a} \rho \hat{a}^\dagger -\frac{\gamma}{2} (\hat{b}^\dagger \hat{b} \rho + \rho \hat{b}^\dagger \hat{b}) + \gamma \hat{b} \rho \hat{b}^\dagger, \nonumber \end{eqnarray} where we have considered $\gamma_{A}=\gamma_{B}=\gamma$ as the local decay rates, $\hat{a}$ and $\hat{b}$ are the lowering operators for the qubits of Alice and Bob respectively, and the interaction picture is implied. As time goes by, entanglement is lost, which can be evidenced by the entanglement of formation $E_F$ of state $\rho(t)$ ~\cite{EfC} as shown in Fig. \ref{fig1}. Due to this process, the further Alice and Bob take to perform the teleportation, the largest the amount of resources needed, since they would need more and more copies in order to distill a maximally entangled pair. Let us now assume that Alice and Bob can monitor the environments of their respective qubits, {\it{i.e.~}} they can continuously detect if their qubits have lost one excitation. It can involve a direct measurement of the reservoir, for example, by detecting an emitted photon, or an indirect probing of the loss of excitation through an auxiliary system~\cite{HarocheNature}. Under these conditions, the time evolution of the system is no longer described by the master equation itself but rather by its so-called quantum-jump unraveling, in which the no-jump operator (corresponding to no detection in the monitored environment) is given by $\hat{\Pi}_{0} = \openone-\frac{dt\gamma}{2}[\hat{a}^\dagger \hat{a}+\hat{b}^\dagger \hat{b}]$ and the one-jump operator is given by $\hat{\Pi}_{1,A}=\sqrt{dt \gamma} \hat{a}$~\cite{RevTrajec}. The same holds for Bob. First, since a detection in the environment immediately kills the entanglement, only the no-jump trajectory is useful. Furthermore, given that $\gamma_A=\gamma_B$, the initial state is preserved under this same evolution, {\it{i.e.~}} $|\Psi(kdt)^{NJ}\rangle=|\Psi(0)\rangle$ for any $k$, keeping its original entanglement. The effect of the reservoir is to make the no-jump trajectory less and less probable with time, its probability given by $P_{NJ}=e^{-\gamma T}$. Provided that Alice and Bob can classically communicate the absence of jumps in their environments, this method is an alternative to the usual distillation protocol since at any time the two parts share a maximally entangled state. \subsection{Optimal singlet conversion through environment monitoring} The idea of locally monitoring the reservoir can also be used in the problem of converting a partially entangled state $\ket{\psi_0}=\sqrt{\alpha}\ket{00}+\sqrt{1-\alpha}\ket{11}$ (with $0\leq\alpha\leq 1/2$) into a maximally entangled one (singlet conversion) \cite{Vidal}. Following a similar reasoning to the 2-qubit protocol presented before, when the subsystems undergo spontaneous decay and no jump is detected in both local reservoirs the initial state $\ket{\psi_0}$ evolves into $\ket{\tilde{\psi}'}=\sqrt{\alpha}\ket{00}+\sqrt{1-\alpha}e^{-\gamma t}\ket{11}$, with probability $\alpha+(1-\alpha)e^{-2\gamma t}$. When $e^{-\gamma t}=\sqrt{\alpha/(1-\alpha)}$, $\ket{\psi'}$ is a maximally entangled state~\cite{Alejo}. This happens with probability $p_{ok}=2\alpha$, which is exactly the optimal singlet conversion probability found in \cite{Vidal}. This shows that indeed coupling the qubits with independent reservoirs and monitoring their environments provides a physical implementation of the optimal singlet conversion protocol. A setup that implements this conversion has recently been done with twin photons \cite{Alejo}. \begin{figure*} \centering \includegraphics[scale=1]{Commu2} \caption{(Color online) Average entanglement over different trajectories, given by $E=\sum_{j_i}E(\psi^{j_i})P^{j_i}$, as a function of $\gamma t$. $E_F$ represents the entanglement evolution without any protection strategy. The numerical index corresponds to the dimension of the system in which both logical qubits are encoded. Index ``f'' indicates the presence of feedback in the system. In $E_{3}$ and $E_{3,f}$ $\gamma_{21} =\gamma_{10}$ (fully degenerated decay channels) whereas index ``h.o.'' indicates decay channels similar to the dissipative harmonic oscillator, i.e. $\gamma_{21} = 2\gamma_{10}$. $\tau$ is the feedback time delay, and $\eta$ is the measurement efficiency. In the right panel we repeat the $E_2$ curve to emphasize that the use of qutrits is advantageous even in the cases with non-perfect measurements or feedback.}\label{fig1} \end{figure*} \section{Encoding qubits in three-level systems} We now proceed to show that using qutrits ($\{|0\rangle,|1\rangle,|2\rangle\}$) instead of qubits in each side improves the previous strategies. Alice and Bob will always share an initial state in the subspace spanned by $\{|1\rangle,|2\rangle\}_A\otimes \{|1\rangle,|2\rangle\}_B$, which will represent our particular encoding. We will also consider, from now on, only cascade decay channels, $|2\rangle \mapsto |1\rangle$ and $|1\rangle \mapsto |0\rangle$ of respective rates $\gamma_{21}$ and $\gamma_{10}$ and frequencies $\omega_{21}$ and $\omega_{10}$. This condition is naturally found in many systems such as harmonic oscillators undergoing dissipation, spontaneous emission of spin-1 systems or some three-level atoms. We begin by analyzing the most favorable situation: the complete degeneracy between the decaying channels in each qutrit ($\omega_{21}=\omega_{10}$ and $\gamma_{21}=\gamma_{10}$), \textit{i.e.} if qutrit $A$ emits an excitation to its reservoir, by detecting this emission there is no fundamental way to identify the corresponding decay channel (both are equally probable and generate indistinguishable excitations). First of all, note that in this case the no-jump evolution does not affect the entire subspace initially used by Alice and Bob, which mimics a decoherence free subspace for this particular trajectory. Furthermore, contrary to the previous case, the detection of one jump does not mean entanglement loss. For example, if Alice and Bob share the initial state $|\Psi(0)\rangle = \frac{|12\rangle+|21\rangle}{\sqrt{2}}$, and Alice detects an excitation in her environment, the state of the system must still be given by $\hat{\Pi}_{1,A}|\Psi(0)\rangle$. But now, the jump operator must include the essential fact that the excitation does not distinguish its \textit{donor}, therefore, it is given by $\sqrt{\gamma dt}(|1\rangle\langle 2| + |0\rangle\langle 1|)$. When applied to $|\Psi(0)\rangle$, this jump operator produces the state $|\Psi^{1,A}(dt)\rangle=\frac{|02\rangle+|11\rangle}{\sqrt{2}}$. This state is still maximally entangled, when interpreted as the state of two logical qubits. Note that the information that the system has decayed not only allows Alice to keep an entangled state with Bob but actually gives her the opportunity to locally and deterministically restore $|\Psi(0)\rangle$ through the unitary feedback of her qutrit with one excitation ($|1\rangle\mapsto |2\rangle$, $|0\rangle\mapsto |1\rangle$). This operation should be fast enough to avoid a consecutive decay which would surely kill any entanglement. Since, at least in principle, this procedure could be repeated over and over again, entanglement could be protected for as long as necessary. The same argument holds for Bob. In fact, the same logic can be applied to a $n$-party multi-qutrit system in which each part is subjected to dissipation but is also able to monitor the loss of excitations and to locally feedback them. Once again, as long as no part loses two excitations in a row (before the feedback mechanism comes in place), an entire subspace spanned by $\{|1\rangle,|2\rangle\}^{\otimes n}$ can be protected against local dissipation. Another possible configuration features emitted excitations that still do not identify the possible decay channel but are more likely the higher the number of excitations in $|\Psi(0)\rangle$ ($\omega_{21}=\omega_{10}$ and $\gamma_{21}>\gamma_{10}$). For example, in the specific case of dissipative harmonic oscillators, the operators in the studied subspace ($|n=0,1,2\rangle$) are given by $\hat{\Pi}_0=\openone-\frac{dt \gamma}{2} (|1\rangle \langle 1| + 2 |2\rangle \langle 2|)$ and $\hat{\Pi}_1= \sqrt{dt \gamma} (|0\rangle \langle 1| + \sqrt{2} |1\rangle \langle 2|)$. In this case, both possible evolutions would affect the entanglement between Alice's and Bob's qutrits, since they redistribute population among the levels of each subsystem. In Fig.~\ref{fig1}a we plot the time evolution of the entanglement of the non-monitored system, $E_F$, as well as the average entanglement over different trajectories~\cite{AndreE} as a function of time for the cases analyzed above, both with and without feedback. Note that in all situations, even for the usual qubit encoding, monitoring the environments preserves entanglement for a longer time than ignoring it, as should be the case since some information on the system is always recovered. When the detection and feedback mechanisms are ideal ($E_{3,f}$), entanglement can be preserved indefinitely. In Fig.~\ref{fig1}b we plot the entanglement obtained either when there is an $8 \%$ delay in the feedback mechanism or when there is an $8 \%$ inefficiency in the measurement of the environment. Since we now deal with mixed states of two qutrits, we use Negativity~\cite{Neg} to obtain the entanglement in each case. We should stress the fact that the use of qutrits is advantageous even in these non-ideal cases when compared to two qubits, as it is clear in Fig.~\ref{fig1}b. Also note that, as expected, inefficient detection of the environment has a greater effect on entanglement loss than delay in the feedback mechanism. \section{Possible implementation using Cavity Eletrodynamics} Let us now describe a cavity QED setup where our ideas can be applied. Suppose Alice and Bob want to establish a (ideally) perfect quantum channel. In the proposal of Ref.~\cite{Luiz}, Alice first prepares an empty cavity (state $\ket{0}$) and then interacts it with a very stable two-level atom ($\{|g\rangle,|e\rangle\}$) in the exited state $\ket{e}$, in such a way to get the entangled state $\sqrt{\alpha}\ket{0e}+\sqrt{1-\alpha}\ket{1g}$. Afterwards she sends the atom to Bob. Assuming that the atomic decay rate is much smaller than the cavity field one, {\it{i.e.~}} $ \gamma_{at} \ll \gamma$, the decoherence time of the second would limit the distance achievable by the first. In fact, the distance between Alice and Bob would have to be much smaller than $v/\gamma$, $v$ being the atomic velocity. As discussed before, this strategy can be improved if Alice uses three levels of her cavity field. The idea is that, instead of being empty, the cavity initially contains a photon (state $\ket{1}$). Then, after the atom-cavity interaction the final state will be $\sqrt{\alpha}\ket{1e}+\sqrt{1-\alpha}\ket{2g}$. If no jump happens in the interval $[0,t)$ then this state evolves to $|\tilde{\chi}_{\mathrm{NJ}}(t)\rangle=e^{-\gamma t/2}(\sqrt{\alpha}|e1\rangle+\sqrt{1-\alpha}e^{-\gamma t/2}|g2\rangle)$, with probability $P_{\chi_0}(t)=\alpha e^{-\gamma t}+(1-\alpha)e^{-2\gamma t}$. However, if one jump happens in the cavity field at time $t_J<t$, then the new state of the system at time $t$ is given by $|\tilde{\chi}_{\mathrm{OJ}}(t)\rangle=e^{-\gamma t_{\mathrm{J}}/2}(\sqrt{\alpha}|e0\rangle+\sqrt{1-\alpha}\sqrt{2}e^{-\gamma t/2}|g1\rangle)$, where the no-jump operator is used for $t < t_J$ and $t > t_J$. Note that because $\gamma_{21}=2\gamma_{10}$, the final state $|\tilde{\chi}_{\mathrm{OJ}}(t)\rangle$ is always the same irrespective of the particular time at which the jump happened. We can then treat all one-jump trajectories as the same and compute their joint probability as $P_{\chi_1}(t)=\int_{0}^{t}P_{\chi_1}(t,t_{\mathrm{J}})dt_{\mathrm{J}}=p\left[\frac{1-e^{-\gamma t}}{\gamma}\right]$, where $P_{\chi_1}(t,t_{\mathrm{J}})=pe^{-\gamma t_{\mathrm{J}}}$ gives the probability density for the trajectory where a jump happens at $t=t_J$ (with $p=\alpha+2(1-\alpha)e^{-\gamma t}$). The trajectory-averaged entanglement is then given by $E_{2\otimes3}(\alpha,t)=E_{\chi_{NJ}}P_{\chi_0}+E_{\chi_{OJ}}P_{\chi_1}$. Furthermore, we can choose $t$ to maximize the one-jump trajectory and $\alpha$ to maximize the entanglement of the state when the atom reaches Bob, hence, creating an even longer lasting entanglement. Fig.~\ref{Superficie2} compares the protocols as a function of $t$ and $\alpha$ showing the optimum range in which each strategy is better and also that for all times and initial states, encoding qubits in qutrits is more efficient in the presence of decay and environmental monitoring. Fig.~\ref{Superficie2}d shows entanglement for different temperatures of the reservoir. Typical microwave Cavity QED experiments correspond to the intermediate curve~\cite{HarocheNature}. Note that at such low temperatures, thermal excitations play a minor role in the proposed scheme. \begin{figure}[h!] \includegraphics[width=8.5cm]{compare}\\ \caption{(Color Online) a. No-jump contribution $E_{\chi_{NJ}}P_{\chi_0}$ for $E_{2\otimes 3}$; b. (upper right) one-jump contribution $E_{\chi_{OJ}}P_{\chi_1}$ for $E_{2 \otimes 3}$; c. $E_{2\otimes3}-E_{2\otimes2}$; d. $E_{2\otimes3}$ with $\alpha=0.5$ for different temperatures with the mean environment excitation number given by $n$. Decay time $1/\gamma=0.129s$ as in the experimental setup in~\cite{HarocheNature}.} \label{Superficie2} \end{figure} \section{Applications for communication protocols} The schemes proposed before can be readily incorporated in some previously known communication protocols. One of the main ideas of transmitting quantum information in 1-D networks relies on quantum repeaters \cite{Repeat}. In this case, one wants to transmit entanglement through large distances and uses intermediate stations to recover it from time to time through entanglement distillation. Once some degree of entanglement is recovered in an intermediate station, this entanglement is teleported to the next station. Naturally, since using qutrits allows to keep entanglement for longer times between each station, there is a reduction in the total amount of necessary resources. In higher dimensional networks, it was recently shown that the geometry of the graph defining the quantum network plays an important role in the problem of long-distance communication. Entanglement Percolation ideas \cite{EntPerc} were first presented for pure-state based networks. The main building block for these strategies is the possibility of performing the optimal singlet conversion. As we discussed before, this task can be achieved by reservoir monitoring. This fact in turn shows a strategy of realizing Classical Entanglement Percolation in noisy lattices \cite{MixedPerc}. \section{conclusion} To conclude, we have shown that for systems undergoing dissipation, encoding qubits in qutrits preserves coherence and entanglement for much longer times if the excitations given to the local reservoirs can be detected by external observers. Thus, systems with their environments under continuous measurement are not only suitable but advantageous for some quantum information protocols. Moreover, the monitoring scheme is further improved if the parties are able to locally feedback the recovered information into the system. This leads naturally to a reduction in resources for protocols such as quantum repeaters, teleportation, swapping and error correction, as well as an improvement in the coherence time of quantum memories. We have also shown a way of using the information leakage to the environment to perform the optimal singlet conversion, the central task in Classical Entanglement Percolation. Finally, note that experiments already observe quantum jumps in different systems, e.g, harmonic oscillators (microwave cavity fields)~\cite{HarocheNature}, and in single ions~\cite{ion}, which means that the scheme here proposed is clearly within nowadays technology. \begin{acknowledgements} The authors thank A.~Ac\'in and A.R.R.~Carvalho for enlightening discussions. Support from Brazilian agencies CNPq and Fapemig and the European project QAP is warmly acknowledged. This work is part of Brazilian National Institute of Science and Technology on Quantum Information. \end{acknowledgements}
\section{Introduction} Let $p$ be a prime, $\chi$ a non-principal character modulo $p$. We denote $e_p(y):=\exp(2\pi iy/p)$ as usual. Sums of the form \begin{equation}\label{eq:definition} \sum_{x=N}^{N+H}\chi(x)e_p(ax), \end{equation} are often encountered in analytic number theory. We call the sums~\eqref{eq:definition} {\it pure} character sums if $a\equiv 0\pmod{p}$, otherwise {\it mixed} character sums. If $H=p$ we say the sums~\eqref{eq:definition} {\it complete}, otherwise {\it incomplete} (or {\it partial} as Burgess used). In the case of $a\not\equiv 0\pmod{p}$ and $H<p$, sums~\eqref{eq:definition} are usually called partial Gaussian sums, which have been well studied by Vinogradov~\cite{V} and Burgess~\cite{B88}. In this paper we try to generalize Burgess' results to arbitrary finite fields. By a well-known generalization of the P\'olya-Vinogradov inequality we have $$ \sum_{x=N}^{N+H}\chi(x)e_p(ax)\ll p^{1/2}\log p. $$ For pure character sums, it was shown by Burgess \cite{B63} that for any positive integer $r$ we have \begin{equation}\label{eq:Burgess_a=0} \sum_{x=N}^{N+H}\chi(x)\ll H^{1-1/r}p^{(r+1)/{4r^2}}\log p. \end{equation} Fifteen years later, by a modification of his method in proving~\eqref{eq:Burgess_a=0}, Burgess~\cite{B88} proved the following estimates for general partial Gaussian sums. \begin{thm}\label{thm:Burgess} Let $\chi$ be a non-principal character modulo a prime $p$. Then for any integers $r\ge 2$, $a$, $N$ and $1\le H < p$ we have \begin{equation}\label{eq:Burgessgeneral} \sum_{x=N}^{N+H}\chi(x)e_p(ax)\ll H^{1-1/r}p^{1/{4(r-1)}}\log^2 p. \end{equation} \end{thm} On the other hand, parallel to the pure character sums~\eqref{eq:Burgess_a=0} in prime field $\F_p$, there are also many works on pure character sums in general finite fields $\F_q$, $q=p^n$. See the papers of Davenport and Lewis~\cite{DL}, Chang~\cite{Ch1} and Konyagin~\cite{Kon}. So it is naturally to consider partial Gaussian sums in arbitrary finite fields. However, such a generalization is quite unusually because the additive character $e_p(\cdot)$ causes additional difficulty even in the case of prime field. Indeed Burgess himself has remarked that {\it the argument used to obtain~\eqref{eq:Burgess_a=0} depended on the summand being multiplicative} (see Burgess~\cite[p. 589]{B88}). Thus the method used by Burgess does not have any natural extensions to the case of arbitrary finite fields. And even nowadays, although the results we obtain in this paper match Burgess' results in the same range, they are not as explicit as those of Burgess. Recently, Chamizo~\cite{C} presented a new proof of Burgess' partial Gaussian sums on the Third Conference on Number Theory at University of Salamanca (Salamanca, July 2009). Chamizo's used essentially the classical {\it method of amplification}\footnote{The method of amplification was first used in number theory by Vinogradov~\cite{Vino}, then introduced by Karatsuba~\cite{K} into the study of character sums. Now it is a classical method, see Friedlander~\cite{F}, Iwaniec and Kowalski~\cite{IK}, Chang~\cite{Ch1}.} in the form of Iwaniec and Kowalski~\cite{IK}. He ingeniously introduced a trick to overcome the difficulty caused by additive character. In the present paper we generalize Burgess' partial Gaussian sums to arbitrary finite fields. Two deep results on multiplicative energy for subsets in finite fields, which are obtained respectively by some tools from additive combinatorics and geometry of numbers, are involved here. We will also use Chamizo's trick. We finally remark that Perel'muter~\cite{P} has studied partial Gaussian sums over additive subgroup of $\F_{p^n}$. However he mainly concerned with the algebraic respects. \section{Notation} Throughout the paper we will use the following notations. Let $p$ be an odd prime, $q$ an integer with $q=p^n$, and $\F_p$ the prime field. Let $\F_q$ denote the finite field with $q$ elements. We recall that the function $$ \Tr(z)=\sum_{i=0}^{n-1}z^{p^i} $$ is called the {\it trace} of $z\in\F_{p^n}$ over $\F_p$. Define $e_p(z)=\exp(2\pi iz/p)$. Then the set of functions $\psi_a(z)=e_p(\Tr(az))$, $a\in\F_{p^n}$, form the set of additive characters of $\F_{p^n}$, with $\psi_0$ being the trivial character. Let $\chi$ be a nontrivial multiplicative character of $\F_{p^n}$. Let $\{\omega_1,\ldots,\omega_n\}$ be an arbitrary basis for $\F_{p^n}$ over $\F_p$. Then the elements of $\F_{p^n}$ have a unique representation as \begin{equation}\label{eq:elements} \xi=x_1\omega_1+\cdots+x_n\omega_n,\qquad 0\le x_i<p. \end{equation} We denote by $B$ a box in the $n$-dimensional space, defined by \begin{equation}\label{eq:box} N_j<x_j\le N_j+H_j,\qquad 1\le j\le n, \end{equation} where $N_j,H_j$ are integers satisfying $0\le N_j<N_j+H_j<p$ for all $j$. For $A\subset\F_q$, we denote by \begin{equation}\label{eq:multiplicetive_energy} E(A):=|\{(x_1,x_2,x_3,x_4)\in A\times A\times A\times A: x_1x_2=x_3x_4\}| \end{equation} the {\it multiplicative energy} of $A$. As usual, `$O$' and `$\ll$' denote respectively Landau and Vinogradov symbol, in which the constants implied depend only on $n$ throughout this paper. \section{Preliminary} \subsection{Pure character sums in finite fields} Davenport and Lewis~\cite{DL} proved in 1963 that \begin{thm}\label{thm:Davenport&Lewis} Let $H_j=H$ for $1\le j\le n$ with $$ H>p^{\frac{n}{2(n+1)}+\eps}\quad\text{for some}~\eps>0, $$ and let $p>p_1(\eps)$, then $$ \left|\sum_{x\in B}\chi(x)\right|<(Hp^{-\delta})^n, $$ where $\delta=\delta(\eps)>0$. \end{thm} \begin{rem} We see that if $n=1$, the exponent in Theorem~\ref{thm:Davenport&Lewis} is still $1/4+\eps$, which recovers Burgess' result. While as $n$ increases, the exponent $\frac{n}{2(n+1)}$ will be near to $1/2$. \end{rem} About two years ago, M.-C.~Chang wrote a series of papers to introduce some tools from additive combinatorics, mainly the sum-product theorems in finite fields, into the study of character sums estimates. She obtained many interesting results, one of which improved Davenport and Lewis~\cite{DL} by combining Burgess' classical amplification method with some estimates for multiplicative energy for subsets in $\F_{p^n}$. \begin{thm}\label{thm:Chang} Let $\chi$ be a nontrivial multiplicative character of $\F_{p^n}$. Given $\eps>0$, there is $\tau>\eps^2/4$ such that if $$ B=\left\{\sum_{j=1}^nx_j\omega_j:x_j\in (N_j,N_j+H_j]\cap\Z,\, 1\le j\le n\right\} $$ is a box satisfying $$ \prod_{j=1}^n H_j>p^{(\frac25+\eps)n} $$ then for $p>p(\eps)$, $$ \left|\sum_{x\in B}\chi(x)\right|\ll_n |B|p^{-\tau}, $$ unless $n$ is even and $\chi\mid_{F_2}$ is principal, where $F_2$ is the subfield of size $p^{n/2}$, in which case, $$ \left|\sum_{x\in B}\chi(x)\right|\le\max_{\xi}|B\cap\xi F_2|+O_n(|B|p^{-\tau}). $$ \end{thm} \begin{rem} Theorem~\ref{thm:Chang} also holds if we replace the assumption $\prod_{j=1}^n H_j>p^{(\frac25+\eps)n}$ by the stronger one $$ H_j > p^{2/5+\eps},\qquad\text{for all}~j, $$ which improved upon Davenport and Lewis~\cite{DL} for $n>4$. But, for higher-dimensional generalization, the results do not achieve the strength of Burgess~\cite{B62}. We note that Burgess' strength is obtained only for some special cases, see Burgess~\cite{B67}, Karatsuba~\cite{K} and Chang~\cite{Ch2}. \end{rem} The main ingredient in Chang~\cite{Ch1} is the following estimate for the multiplicative energy. \begin{prop}\label{prop:ChangEnergy} Let $\{\omega_1,\ldots,\omega_n\}$ be a basis for $\F_{p^n}$ over $\F_p$, and let $B\subset\F_{p^n}$ be the box $$ B=\left\{\sum_{j=1}^nx_j\omega_j:x_j\in[N_j+1,N_j+H_j],\, j=1,\ldots,n\right\} $$ where $1\le N_j<N_j+H_j<p$ for all $j$. Assume that \begin{equation}\label{eq:condition} \max_j H_j<\frac12(\sqrt{p}-1). \end{equation} Then we have $$ E(B,B)<C^n(\log p)|B|^{11/4} $$ for an absolute constant $C<2^{9/4}$. \end{prop} \begin{rem} Using a result of Perel'muter and Shparlinski~\cite{PS} and some sophisticated arguments, Chang removed the influence of the condition~\eqref{eq:condition} on Theorem~\ref{thm:Chang}. \end{rem} On the conference of 26th Journ\'ees Arithm\'etiques (Saint-Etienne, July 2009), using the method in geometry of numbers (see~\cite{Ban}, \cite{TV}), Konyagin~\cite{Kon} improved Chang's estimate for multiplicative energy if $H_i=H$, $1\le i\le n$. \begin{prop}\label{prop:KonyaginEnergy} If $H_1=\cdots=H_n\le p^{1/2}$, then $$ E(B)\ll |B|^2\log p. $$ \end{prop} Then, incorporating the estimate of Proposition~\ref{prop:KonyaginEnergy} into Burgess' amplification process, Konyagin proved \begin{thm}\label{thm:Konyagin} Let $\chi$ be a nontrivial multiplicative character of $\F_{p^n}$ and $0<\eps\le 1/4$ be given. If $n\ge 2$ and $B$ is a box defined in~\eqref{eq:box} and satisfying $$ H_j\ge p^{1/4+\eps},\qquad 1\le j\le n, $$ then $$ \left|\sum_{x\in B}\chi(x)\right|\ll |B|p^{-\eps^2/2}. $$ \end{thm} \subsection{Weil's theorem} We will need the following version of Weil's bound on exponential sums. See~\cite[Theorem 11.23]{IK}. \begin{thm}[A. Weil]\label{thm:Weil} Let $\chi$ be a nontrivial multiplicative character of $\F_{p^n}$ of order $d>1$. Suppose $f\in\F_{p^n}[x]$ has $m$ distinct roots and $f$ is not a $d$-th power. Then for $n\ge 1$ we have $$ \left|\sum_{x\in\F_{p^n}}\chi(f(x))\right| \le (m-1)p^{\frac{n}{2}}. $$ \end{thm} \section{Main results} The following two theorems generalize Theorem~\ref{thm:Chang} and Theorem~\ref{thm:Konyagin} respectively. \begin{thm}\label{thm:CG} Let $\chi$ be a nontrivial multiplicative character of $\F_{p^n}$. Given $\eps>0$, there is $\tau>\eps^2/4$ such that if $B$ is a box defined in~\eqref{eq:box} and satisfying $$ \prod_{j=1}^n H_j \ge p^{\(\frac25+\eps\)n}, $$ then for $p>p(\eps)$, $$ \left|\sum_{x\in B}\chi(x)e_p(\Tr(ax))\right|\ll |B|p^{-\tau}, $$ unless $n$ is even and $\chi|_{F_2}$ is principal, where $F_2$ is the subfield of size $p^{n/2}$, in which case, $$ \left|\sum_{x\in B}\chi(x)e_p(\Tr(ax))\right|\le\max_{\xi}|B\cap\xi F_2| + O_n(p^{-\tau}|B|). $$ \end{thm} \begin{thm}\label{thm:KG} Let $\chi$ be a nontrivial multiplicative character of $\F_{p^n}$ and $0<\eps\le 1/4$ be given. If $n\ge 2$ and $B$ is a box defined in~\eqref{eq:box} and satisfying $$ H_j\ge p^{1/4+\eps},\qquad 1\le j\le n, $$ then $$ \left|\sum_{x\in B}\chi(x)e_p(\Tr(ax))\right|\ll |B|p^{-\eps^2/2}. $$ \end{thm} \section{Proof of Theorem~\ref{thm:CG}} We incorporate the technique of Chamizo~\cite{C} into the argument of Chang~\cite{Ch1}. \begin{proof}[Proof of Theorem~\ref{thm:CG}] We first prove the theorem under the restriction \begin{equation}\label{eq:H_j} H_j<\frac12(\sqrt{p}-1)\qquad\text{for all}~j, \end{equation} which is inherited from the estimate for multiplicative energy in Proposition~\ref{prop:ChangEnergy}. By breaking up $B$ in smaller boxes, we may assume \begin{equation}\label{eq:Chang_H} \prod_{j=1}^n H_j\sim p^{(\frac25+\eps)n}. \end{equation} Let $\delta>0$ be specified later. Let $$ I=[1,p^{\delta}], $$ $$ B_0=\left\{\sum_{j=1}^n x_j\omega_j : x_j\in[0,p^{-2\delta}H_j],\,j=1,\ldots,n \right\} $$ and $$ B_I=\left\{\sum_{j=1}^n x_j\omega_j : x_j\in[0, p/|I|^{\frac{1}{n}}],\, j=1,\ldots,n \right\}. $$ Since $B_0 I\subset\left\{\sum_{j=1}^n x_j\omega_j : x_j\in[0, p^{-\delta}H_j],\, j=1,\ldots,n \right\}$, clearly \begin{eqnarray*} \lefteqn{\left|\sum_{x\in B}\chi(x)e_p(\Tr(ax)) - \sum_{x\in B}\chi(x+yz)e_p\(\Tr(a(x+yz))\)\right|} \\ & & \quad < |B\setminus(B+yz)| + |(B+yz)\setminus B| < 2np^{-\delta}|B| \end{eqnarray*} for $y\in B_0$, $z\in I$. Hence \begin{eqnarray*} \lefteqn{\sum_{x\in B}\chi(x)e_p(\Tr(ax))} \\ & & \qquad =\frac{1}{|B_0|\,|I|}\sum_{x\in B,\,y\in B_0,\,z\in I}\chi(x+yz)e_p\(\Tr(a(x+yz))\)+O(np^{-\delta}|B|). \end{eqnarray*} We now estimate \begin{eqnarray*} & & \left|\sum_{x\in B,\,y\in B_0,\,z\in I}\chi(x+yz)e_p\(\Tr(a(x+yz))\)\right| \\ & \le & \sum_{x\in B,\,y\in B_0}\bigg|\sum_{z\in I}\chi(x+yz)e_p\(\Tr(a(x+yz))\)\bigg| \\ & \le & \sum_{u\in\F_q}\nu(u)\sup_{y\in B_0}\left|\sum_{z\in I}\chi(u+z)e_p\(\Tr(ay(u+z))\)\right| \\ & \le & \sum_{u\in\F_q}\nu(u)\sup_b\left|\sum_{z\in I}\chi(u+z)e_p\(\Tr(b(u+z))\)\right|, \end{eqnarray*} where $$ \nu(u)=\Big|\Big\{(x,y)\in B\times B_0:\frac{x}{y}=u\Big\}\Big|. $$ Then \begin{eqnarray*} \lefteqn{\sum_{x\in B}\chi(x)e_p(\Tr(ax))} \\ & & \qquad \le\frac{1}{|B_0|\,|I|}\sum_{u\in\F_q}\nu(u)\sup_b\left|\sum_{z\in I}\chi(u+z)e_p\(\Tr(b(u+z))\)\right| + O(np^{-\delta}|B|) \\ & & \qquad \le\frac{1}{|B_0|\,|I|}\sum_{u\in\F_q}\nu(u)\sup_b\frac{|I|}{q}\sum_{\substack{c\\ c-b\in B_I}}\left|\sum_{z\in I}\chi(u+z)e_p\(\Tr(c(u+z))\)\right| + O(np^{-\delta}|B|) \\ % & & \qquad \le\frac{1}{|B_0|\,q}\sum_{u\in\F_q}\nu(u)\sum_{\substack{c\\ c-b_0\in B_I}}\left|\sum_{z\in I}\chi(u+z)e_p\(\Tr(c(u+z))\)\right| + O(np^{-\delta}|B|) % \end{eqnarray*} with the sum $$ \sum_{\substack{c\\ c-b\in B_I}}\left|\sum_{z\in I}\chi(u+z)e_p\(\Tr(c(u+z))\)\right| $$ attains its maximum at $b_0\in\F_q$. Let $r\ge 2$ by any integer. Applying the H\"older inequality $$ \sum_{u\in\F_q}\nu(u)\sum_{\substack{c\\ c-b_0\in B_I}}\left|\sum_{z\in I}\chi(u+z)e_p\(\Tr(c(u+z))\)\right|\le V_1^{1-\frac{1}{r}} V_2^{\frac{1}{2r}} W^{\frac{1}{2r}}, % $$ where $$ V_1=\sum_{u\in\F_q}\nu(u),\quad V_2=\sum_{u\in\F_q}\nu^2(u), $$ $$ W=\sum_{u\in\F_q}\(\sum_{\substack{c\\ c-b_0\in B_I}}\left|\sum_{z\in I}\chi(u+z)e_p\(\Tr(c(u+z))\)\right|\)^{2r}. % $$ Observe that $$ V_1 = |B||B_0| $$ and \begin{eqnarray*} V_2 % & = & |\{(x_1,x_2,y_1,y_2)\in B\times B\times B_0\times B_0:x_1y_2=x_2y_1\}| \\ & = & \sum_v|\{(x_1,x_2):\frac{x_1}{x_2}=v\}|\,|\{(y_1,y_2):\frac{y_1}{y_2}=v\}| \\ & \le & E(B,B)^{\frac12}E(B_0,B_0)^{\frac12} \\ & < & 2^{\frac94n+1}(\log p)|B|^{\frac{11}{8}}|B_0|^{\frac{11}{8}} \\ & < & 2^{\frac94n+1}(\log p)|B|^{\frac{11}{4}}p^{-\frac{11}{4}n\delta}, \end{eqnarray*} by the Cauchy-Schwarz inequality, Proposition~\ref{prop:ChangEnergy} and the definition of $B_0$. Now we bound $W$. Recall that $$ q=p^n. $$ Then \begin{eqnarray*} \lefteqn{\sum_{u\in\F_q}\(\sum_{\substack{c\\ c-b_0\in B_I}}\left|\sum_{z\in I}\chi(u+z)e_p\(\Tr(c(u+z))\)\right|\)^{2r}} \\ & & \le (q/|I|)^{2r-1}\sum_{u\in\F_q}\sum_{c\in\F_q}\left|\sum_{z\in I}\chi(u+z)e_p\(\Tr(c(u+z))\)\right|^{2r} \\ & & \le (q/|I|)^{2r-1}\sum_{z_1,\,\ldots,\,z_{2r}\in I}\bigg|\sum_{u\in\F_q}\chi((u+z_1)\cdots(u+z_r)(u+z_{r+1})^{q-2}\cdots(u+z_{2r})^{q-2}) \\ & & \qquad \times\sum_{c\in\F_q}e_p\(\Tr(c(z_1+\cdots+z_r-z_{r+1}-\cdots-z_{2r}))\)\bigg| \\ & & = q^{2r}|I|^{1-2r}\sum_{\substack{z_1,\,\ldots,\,z_{2r}\in I\\ z_1+\cdots+z_r=z_{r+1}+\cdots+z_{2r}}} % \bigg|\sum_{u\in\F_q}\chi((u+z_1)\cdots(u+z_r)(u+z_{r+1})^{q-2}\cdots(u+z_{2r})^{q-2})\bigg|. \end{eqnarray*} The last equality is by the orthogonality of additive characters. For $z_1,\ldots,z_{2r}\in I$ such that at least one of the elements is not repeated twice, the polynomial $f_{z_1,\ldots,z_{2r}}(u)=(u+z_1)\cdots(u+z_r)(u+z_{r+1})^{q-2}\cdots(u+z_{2r})^{q-2}$ clearly cannot be a $d$-th power. Since $f_{z_1,\ldots,z_{2r}}(u)$ has no more than $2r$ many distinct roots, Theorem~\ref{thm:Weil} gives $$ \bigg|\sum_{u\in\F_q}\chi((u+z_1)\cdots(u+z_r)(u+z_{r+1})^{q-2}\cdots(u+z_{2r})^{q-2})\bigg|<2rp^{\frac{n}{2}}. $$ For those $z_1,\ldots,z_{2r}\in I$ such that every root of $f_{z_1,\ldots,z_{2r}}(u)$ appears at least twice, we bound $\sum\bigl|\sum_{u\in\F_q}\chi(f_{z_1,\ldots,z_{2r}}(u))\bigr|$ by $q$ times the number of such $z_1,\ldots,z_{2r}$. Since there are at most $r$ roots in $I$ and for each $z_1,\ldots,z_{2r}$ there are at most $r$ choices, we obtain a bound $|I|^r r^{2r}p^n$. Therefore $$ \sum_{u\in\F_q}\(\sum_{c\in\F_q}\left|\sum_{z\in I}\chi(u+z)e_p\(\Tr(c(u+z))\)\right|\)^{2r}\le q^{2r}|I|(r^{2r}|I|^{-r} p^n+2rp^{\frac{n}{2}}) $$ and $$ W^{\frac{1}{2r}} % \le q|I|^{\frac{1}{2r}}(r|I|^{-\frac12}p^{\frac{n}{2r}}+2p^{\frac{n}{4r}}). $$ Putting the above estimates together, we have \begin{eqnarray*} \frac{1}{|B_0|q}V % & < & 4^{\frac{n}{r}}(\log p)(|B_0||B|)^{-\frac{1}{r}}|B|^{1+\frac{11}{8r}}p^{-\frac{11}{8}\frac{n}{r}\delta} |I|^{\frac{1}{2r}}\Big(r|I|^{-\frac12}p^{\frac{n}{2r}}+2p^{\frac{n}{4r}}\Big) \\ & < & 4^{\frac{n}{r}}(\log p)p^{2\frac{n}{r}\delta-\frac{11}{8}\frac{n}{r}\delta}|B|^{1-\frac{5}{8r}} |I|^{\frac{1}{2r}}\Big(r|I|^{-\frac12}p^{\frac{n}{2r}}+2p^{\frac{n}{4r}}\Big) \\ & < & 2r\cdot4^{\frac{n}{r}}|B|^{1-\frac{5}{8r}}p^{\frac{n}{4r}+\frac58\frac{n}{r}\delta+\frac{\delta}{2r}}\log p \\ & < & 2r\cdot4^{\frac{n}{r}}|B|p^{\frac{n}{4r}+\frac58\frac{n}{r}\delta-\frac58\frac{n}{r}(\frac25+\eps)+\frac{\delta}{2r}}\log p \\ & < & 2r\cdot4^{\frac{n}{r}}|B|p^{-\frac58\frac{n}{r}(\eps-\delta)+\frac{\delta}{2r}}\log p. \end{eqnarray*} The second-to-last inequality holds because of \eqref{eq:Chang_H} and by assuming $\delta\ge\frac{n}{2r}$. Similar to the argument of Chang~\cite{Ch1}, we can show that $$ p^{-\frac58\frac{n}{r}(\eps-\delta)+\frac{\delta}{2r}}\log p < p^{-\eps^2/4}. $$ Then we prove the theorem under the condition~\eqref{eq:H_j}. \bigskip Now we are at the position to remove the additional hypothesis~\eqref{eq:H_j} on the shape of $B$. We proceed in several steps and rely essentially on a further key ingredient provided by the following estimate in Perel'muter and Shparlinski~\cite{PS}. \begin{prop}\label{prop:PS} Let $\chi$ be a nonprincipal multiplicative character of $\F_q$ and let $g\in\F_q$ be a generating element, i.e. $\F_q=\F_p(g)$. Then for any $a\in\F_p$, we have \begin{equation}\label{eq:PS} \left|\sum_{t\in\F_p}\chi(g+t)e_p(at)\right|\le np^{1/2}. \end{equation} \end{prop} First we make the following observation. Let $H_1\ge H_2\ge\ldots\ge H_n$. If $H_1<p^{\frac12+\frac{\varepsilon}{2}}$, we may clearly write $B$ as a disjoint union of boxes $B_{\alpha}\subset B$ satisfying the first condition in~\eqref{eq:H_j} and $|B_{\alpha}|>(\frac12 p^{-\frac{\varepsilon}{2}})^n|B|>2^{-n}p^{(\frac25+\frac{\varepsilon}{2})n}$. Since~\eqref{eq:H_j} holds for each $B_{\alpha}$, we have $$ \left|\sum_{x\in B_{\alpha}}\chi(x)e_p(\Tr(ax))\right|<cnp^{-\tau}|B_{\alpha}|. $$ Hence $$ \left|\sum_{x\in B}\chi(x)e_p(\Tr(ax))\right|<cnp^{-\tau}|B|. $$ Therefore we may assume that $H_1>p^{\frac12+\frac{\varepsilon}{2}}$. \bigskip \noindent{\it Case 1. $n$ is odd.} \bigskip We denote $I_i=[N_i+1,N_i+H_i]$. Then \begin{eqnarray*} \lefteqn{\left|\sum_{x\in B}\chi(x)e_p(\Tr(ax))\right|} \\ & & = \left|\sum_{\substack{x_i\in I_i\\ 2\le i\le n}}\sum_{x_1\in I_1}\chi\(x_1+x_2\frac{\omega_2}{\omega_1}+\cdots+x_n\frac{\omega_n}{\omega_1}\) % e_p(\Tr(ax_1\omega_1+x_2\omega_2+\cdots+x_n\omega_n))\right| \\ & & \le \left|\sum_{(x_2,\ldots,x_n)\in D^c}\sum_{x_1\in I_1}\chi\(x_1+x_2\frac{\omega_2}{\omega_1}+\cdots+x_n\frac{\omega_n}{\omega_1}\) % e_p(\Tr(ax_1\omega_1+x_2\omega_2+\cdots+x_n\omega_n))\right| \\ & & \qquad + \left|\sum_{(x_2,\ldots,x_n)\in D}\sum_{x_1\in I_1}\chi\(x_1+x_2\frac{\omega_2}{\omega_1}+\cdots+x_n\frac{\omega_n}{\omega_1}\) % e_p(\Tr(ax_1\omega_1+x_2\omega_2+\cdots+x_n\omega_n))\right| \end{eqnarray*} with $$ D = \left\{(x_2,\ldots,x_n)\in I_2\times\cdots\times I_n : % \F_p\(x_2\frac{\omega_2}{\omega_1}+\cdots+x_n\frac{\omega_n}{\omega_1}\)\neq\F_q\right\} $$ and $$ D^c = I_2\times\cdots\times I_n\setminus D. $$ Using~\eqref{eq:PS} we estimate the first sum as \begin{eqnarray*} \lefteqn{\left|\sum_{(x_2,\ldots,x_n)\in D^c}\sum_{x_1\in I_1}\chi\(x_1+x_2\frac{\omega_2}{\omega_1}+\cdots+x_n\frac{\omega_n}{\omega_1}\) % e_p(\Tr(ax_1\omega_1+x_2\omega_2+\cdots+x_n\omega_n))\right|} \\ & & = \left|\sum_{\substack{x_i\in I_i\\ 2\le i\le n}}e_p(\Tr(a(x_2\omega_2+\cdots+x_n\omega_n))) % \sum_{x_1\in I_1}\chi\(x_1+x_2\frac{\omega_2}{\omega_1}+\cdots+x_n\frac{\omega_n}{\omega_1}\)e_p(\Tr(ax_1\omega_1))\right| \\ & & \le \sum_{\substack{x_i\in I_i\\ 2\le i\le n}} % \left|\sum_{x_1\in I_1}\chi\(x_1+x_2\frac{\omega_2}{\omega_1}+\cdots+x_n\frac{\omega_n}{\omega_1}\)e_p(\Tr(a\omega_1)x_1)\right| \\ & & = \sum_{\substack{x_i\in I_i\\ 2\le i\le n}} % \left|\sum_{x_1\in\F_p}\chi\(x_1+x_2\frac{\omega_2}{\omega_1}+\cdots+x_n\frac{\omega_n}{\omega_1}\)e_p(\Tr(a\omega_1)x_1) % \cdot\frac{1}{p}\sum_{b\in\F_p}\sum_{x'_1\in I_1}e_p(b(x_1-x'_1))\right| \\ & & \le \frac{1}{p}\sum_{\substack{x_i\in I_i\\ 2\le i\le n}}\sum_{b\in\F_p} % \left|\sum_{x_1\in\F_p}\chi\(x_1+x_2\frac{\omega_2}{\omega_1}+\cdots+x_n\frac{\omega_n}{\omega_1}\)e_p((\Tr(a\omega_1)+b)x_1)\right| % \cdot\left|\sum_{x'_1\in I_1}e_p(-bx'_1)\right| \\ & & \le \frac{1}{p}np^{1/2}\frac{|B|}{H_1}\sum_{b\in\F_p}\left|\sum_{x'_1\in I_1}e_p(bx'_1)\right| \\ & & \le c(n)p^{\frac12}\log p\frac{|B|}{H_1}. \end{eqnarray*} For the second sum, we have \begin{eqnarray*} & & \left|\sum_{(x_2,\ldots,x_n)\in D}\sum_{x_1\in I_1}\chi\(x_1+x_2\frac{\omega_2}{\omega_1}+\cdots+x_n\frac{\omega_n}{\omega_1}\) % e_p(\Tr(ax_1\omega_1+x_2\omega_2+\cdots+x_n\omega_n))\right| \\ & & \qquad \le p|D|\le p\sum_G\left|G\bigcap\operatorname{Span}_{\F_p}\(\frac{\omega_2}{\omega_1},\ldots,\frac{\omega_n}{\omega_1}\)\right|, \end{eqnarray*} where $G$ runs over nontrivial subfields of $\F_q$. Since $q=p^n$ and $n$ is odd, obviously $[\F_q:G]\ge 3$. Hence $[G:\F_p]\le\frac{n}{3}$. Furthermore, since $\{\omega_1,\ldots,\omega_n\}$ is a basis of $\F_q$ over $\F_p$, $1\notin\operatorname{Span}_{\F_p}(\frac{\omega_2}{\omega_1},\ldots,\frac{\omega_n}{\omega_1})$ and the proceeding implies that $$ \operatorname{dim}_{\F_p}\(G\bigcap\operatorname{Span}_{\F_p} \(\frac{\omega_2}{\omega_1},\ldots,\frac{\omega_n}{\omega_1}\)\)\le\frac{n}{3}-1. $$ Therefore, under our assumption on $|H_1|$, we have $$ \left|\sum_{x\in B}\chi(x)e_p(\Tr(ax))\right|\le c(n)((\log p)p^{-\frac{\varepsilon}{2}}|B|+p^{\frac{n}{3}}) % < \left(c(n)(\log p)p^{-\frac{\varepsilon}{2}}+p^{-\frac{n}{15}}\right)|B|, $$ since $|B|>p^{\frac25 n}$. This proves our claim. \bigskip \noindent{\it Case 2. $n$ is even.} \bigskip In view of the earlier discussion, we have \begin{eqnarray*} \lefteqn{\left|\sum_{x\in B}\chi(x)e_p(\Tr(ax))\right|} \\ & & \le \left|\sum_{(x_2,\ldots,x_n)\in D_2^c}\sum_{x_1\in I_1}\chi\(x_1+x_2\frac{\omega_2}{\omega_1}+\cdots+x_n\frac{\omega_n}{\omega_1}\) % e_p(\Tr(ax_1\omega_1+x_2\omega_2+\cdots+x_n\omega_n))\right| \\ & & \quad + \left|\sum_{(x_2,\ldots,x_n)\in D_2}\sum_{x_1\in I_1}\chi\(x_1+x_2\frac{\omega_2}{\omega_1}+\cdots+x_n\frac{\omega_n}{\omega_1}\) % e_p(\Tr(ax_1\omega_1+x_2\omega_2+\cdots+x_n\omega_n))\right|, \end{eqnarray*} where $$ D_2 = \left\{(x_2,\ldots,x_n)\in I_2\times\cdots\times I_n : \(x_2\frac{\omega_2}{\omega_1}+\cdots+x_n\frac{\omega_n}{\omega_1}\)\in F_2\right\}, $$ $$ D_2^c = I_2\times\cdots\times I_n\setminus D_2, $$ and $F_2$ is the subfield of size $p^{n/2}$. Our only concern is to bound the second sum, namely $$ \varpi=\left|\sum_{(x_2,\ldots,x_n)\in D_2}\sum_{x_1\in I_1}\chi\(x_1+x_2\frac{\omega_2}{\omega_1}+\cdots+x_n\frac{\omega_n}{\omega_1}\) % e_p(\Tr(ax_1\omega_1+x_2\omega_2+\cdots+x_n\omega_n))\right|. $$ First, we note that since $1,\frac{\omega_2}{\omega_1},\ldots,\frac{\omega_n}{\omega_1}$ are independent, $\frac{\omega_j}{\omega_1}\in F_2$ for at most $\frac{n}{2}-1$ many $j$'s. After reordering, we may assume that $\frac{\omega_j}{\omega_1}\in F_2$ for $2\le j\le k$ and $\frac{\omega_j}{\omega_1}\notin F_2$ for $k+1\le j\le n$, where $k\le\frac{n}{2}$. we also assume that $H_{k+1}\le\cdots\le H_n$. Fix $x_2,\ldots,x_{n-1}$. Obviously there is no more than one value of $x_n$ such that $x_2\frac{\omega_2}{\omega_1}+\cdots+x_n\frac{\omega_n}{\omega_1}\in F_2$, since otherwise $(x_n-x'_n)\frac{\omega_n}{\omega_1}\in F_2$ with $x_n\neq x'_n$ contradicting the fact that $\frac{\omega_n}{\omega_1}\notin F_2$. Therefore, $$ |D_2|\le|I_2|\cdots|I_{n-1}| $$ and $$ \varpi\le\frac{|B|}{H_n}. $$ If $H_n>p^{\tau}$, we are done. Otherwise \begin{equation}\label{eq:H_{k+1}} H_{k+1}\cdots H_n\le p^{(n-k)\tau} < p^{\tau}. \end{equation} Define \begin{equation}\label{eq:B_2} B_2=\left\{x_1+x_2\frac{\omega_2}{\omega_1}+\cdots+x_k\frac{\omega_k}{\omega_1} : x_i\in I_i,\,1\le i\le k\right\}. \end{equation} Hence $B_2\subset F_2$ and by~\eqref{eq:H_{k+1}} $$ |B_2|>\frac{|B|}{H_{k+1}\cdots H_n}>p^{\frac25-\frac{\tau}{2}n}>p^{\frac{n}{3}}. $$ (We assume $\tau<\frac{2}{15}$.) Clearly, if $(x_2,\ldots,x_n)\in D_2$, then $z=x_{k+1}\frac{\omega_{k+1}}{\omega_1}+\cdots+x_n\frac{\omega_n}{\omega_1}\in F_2$. Assume $\chi|_{F_2}$ non-principal. Then by completing the sum over $y$ and recalling the classical estimates for Gaussian sums in finite fields~\cite[Theorem 5.11]{LN} and~\eqref{eq:B_2}, we have $$ \left|\sum_{y\in B_2}\chi(y+z)e_p(\Tr(a\omega_1(y+z)))\right|\le(\log p)^{\frac{n}{2}}\max_{\psi} % \left|\sum_{x\in F_2}\psi(x)\chi(x)\right|\le(\log p)^{\frac{n}{2}}|F_2|^{\frac12}\le p^{-\frac{n}{15}}|B_2|, $$ where $\psi$ runs over all additive characters. Therefore, clearly $$ \varpi\le H_{k+1}\cdots H_n p^{-\frac{n}{15}}|B_2|=p^{-\frac{n}{15}}|B| $$ providing the required estimate. If $\chi|_{F_2}$ is principal, then obviously $$ \varpi = \left|\sum_{x_1\in I_1}\sum_{(x_2,\ldots,x_n)\in D_2} % e_p(\Tr(a(x_1\omega_1+\cdots+x_n\omega_n)))\right| \le H_1 |D_2| = \left|F_2\cap\frac{1}{\omega_1}B\right| $$ and $$ \left|\sum_{x\in B}\chi(x)e_p(\Tr(ax))\right| = |\omega_1F_2\cap B| + O_n(p^{-\tau}|B|). $$ This completes the proof of Theorem~\ref{thm:CG}. \end{proof} \section{Proof of Theorem~\ref{thm:KG}} Similar to the proof of Theorem~\ref{thm:CG}, by breaking up $B$ in smaller boxes, we may assume $$ H_1\asymp\cdots\asymp H_n\asymp p^{\frac14+\eps}. $$ Then, using the arguments in the proof of Theorem~\ref{thm:CG}, we can prove Theorem~\ref{thm:KG} along the lines of Konyagin~\cite{Kon}. \section*{Acknowledgements} The author thanks Professor Igor Shparlinski for his helpful comments on an earlier version of the paper. The author also thanks Professor Sergei Konyagin for sending his preprint. Part of this work was done while the author was visiting the Morningside Center of Mathematics, whose hospitality is gratefully acknowledged. The author was supported by the National Natural Science Foundation of China (Grant No. 10671056). \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
\section{Self-Consistent T-matrix approximation}\label{app.SCTM} The self-consistent T-matrix approximation\cite{ShibaPTP68} is defined by the coupled equations for the average Green's function $\hat{G}_{\alpha\beta }(\mathbf{k})\equiv\delta_{\alpha\beta}\hat{G}_{\alpha}(\mathbf{k})$ and the self-energy part $\hat{\Sigma}_{\alpha}\equiv\delta_{\alpha\beta}\hat{\Sigma }_{\alpha\alpha}$, \begin{subequations} \label{SCTM} \begin{align} & \hat{G}_{\alpha}(\mathbf{k})=\left[ E-\hat{\varepsilon}_{\alpha}\hat{\tau }_{z}+\Delta_{\alpha}\hat{\tau}_{x}-\hat{\Sigma}_{\alpha}\right] ^{-1},\\ & \hat{\Sigma}_{\alpha}=n_{i}\hat{T}_{\alpha}=n_{i}\left\langle \left( 1-\hat{\tau}_{z}U_{\alpha\beta}\hat{G}_{0,\beta}\right) ^{-1}\hat{\tau} _{z}U_{\beta\alpha}\right\rangle \end{align} with $\hat{G}_{0,\alpha}=\int\frac{d^{d}\mathbf{k}}{(2\pi)^{d}}\hat{G} _{\alpha}(\mathbf{k})$. Note that the off-diagonal components of $\hat{\Sigma }_{\alpha\beta}$ average to zero due to the random signs of the off-diagonal components of $U_{\alpha\beta}$. The partial density of states is related to the Green's function as $N_{\alpha}(E)=\operatorname{Im}\left\{ \mathrm{Tr}\left[ \hat{G}_{0,\alpha}(E-i\delta)\right] \right\} $. Using the expansion $\hat{\Sigma}_{\alpha}=\sum_{i}\Sigma_{\alpha,i}\hat{\tau}_{i}$ with $i=0,x,y,z$, we can express $N_{\alpha}(E)$ via the ratio $u_{\alpha}=\left( E-\Sigma_{\alpha,0}\right) /\left( \Delta-\Sigma_{\alpha,x}\right) $ as \end{subequations} \begin{equation} N_{\alpha}(E)=2\nu_{\alpha}\operatorname{Im}\frac{u_{\alpha}}{\sqrt {1-u_{\alpha}^{2}}}.\label{DoS} \end{equation} Here the factor $2$ accounts for the electron and hole excitations. In the case of identical bands, this system reduces to an equation for only one complex parameter $u=\left( E-\Sigma_{0}\right) /\left( \Delta-\Sigma _{x}\right) $, similar to the Shiba equation for magnetic impurities\cite{ShibaPTP68}, \begin{align} & u\left( 1-\alpha\frac{\sqrt{1-u^{2}}}{\varepsilon_{0}^{2}-u^{2}}\right) =\frac{E}{\Delta},\label{ShibaEq}\\ & \alpha=\frac{2n_{i}\Gamma_{\mathrm{eff}}}{\pi\nu\Delta}=\frac{1}{\tau _{12}\Delta}\\ & =\frac{2n_{i}}{\pi\nu\Delta}\frac{\gamma_{12}\gamma_{21}}{1+\gamma_{22} ^{2}+\gamma_{11}^{2}+2\gamma_{12}\gamma_{21}+\left( \gamma_{22}\gamma _{11}-\gamma_{12}\gamma_{21}\right) ^{2}}.\nonumber \end{align} We observe that increasing the intraband scattering potential $\gamma _{\alpha\alpha}$ increases the intraband scattering time $\tau_{12}$ and diminishes the pair-breaking parameter $\alpha$.\cite{SengaJPSJ08} We also note that the SCTM results do not depend on dimensionality of the superconductor. We summarize the most important analytical results obtained within this approximation. The SCTM approximation gives a gapped state for $\alpha <\varepsilon_{0}^{2}$ and gapless state for $\alpha>\varepsilon_{0}^{2}$ with finite total density of state at zero energy \begin{equation} N_{s}(0)=4\nu\frac{\sqrt{\alpha^{2}-\varepsilon_{0}^{4}}}{\sqrt{\frac {\alpha^{2} }{2}+\varepsilon_{0}^{2}\left( 1-\varepsilon_{0}^{2}\right) +\alpha\sqrt{\frac{\alpha^{2}}{4}+1-\varepsilon_{0}^{2}}}}.\label{N0Tmatrix} \end{equation} We stress again that the existence of a \emph{gapped} state is unrealistic feature and deficiency of this approximation. The suppression of $T_{c}$ is determined by the Abrikosov-Gor'kov formula \cite{AGZhETF60}, \begin{equation} \ln\frac{T_{c0}}{T_{c}} =\psi(1/2+1/2\pi\tau_{12} T_{c})-\psi (1/2),\label{SupprTc} \end{equation} where $\psi(x)$ is the digamma function. This famous result is almost always understood too literally. In fact, it just gives an approximate typical value of the critical temperature. In real systems, due to the random arrangement of impurities, the transition temperature is inhomogeneous and the transition has percolative nature. The average zero-temperature gap parameter is determined by the equations \cite{RusinovPZheTF68} \begin{subequations} \begin{equation} \ln\frac{\Delta_{0}}{\Delta}=\frac{\pi}{2}\frac{\alpha}{1+\varepsilon_{0} }\text{ for }\alpha<\varepsilon_{0}^{2}\label{GapSCTM1} \end{equation} with $\alpha=\alpha_{0}\Delta_{0}/\Delta$ and \begin{align} \ln\frac{\Delta_{0}}{\Delta} & =\frac{\pi}{2}\frac{\alpha}{\varepsilon_{0} +1}+\ln\left(u_{0}+\sqrt{1+u_{0}^{2}}\right) -\frac{\alpha u_{0}}{u_{0}^{2}\!+\!\varepsilon_{0}^{2}} \nonumber\\ & \!+\!\frac{\alpha }{\varepsilon_{0}^{2}\!-\!1}\left[ \arctan u_{0}\!-\!\varepsilon_{0} \arctan\left( \frac{u_{0}}{\varepsilon_{0}}\right) \right] ,\label{GapSCTM2}\\ \text{with }u_{0}^{2} =&\frac{\alpha^{2}}{2}-\varepsilon_{0}^{2}+\alpha \sqrt{\frac{\alpha^{2}}{4}+1-\varepsilon_{0}^{2}}, \text{ for } \alpha>\varepsilon_{0}^{2}\nonumber \end{align} \end{subequations} Here $\Delta_{0}$ is the gap parameter for the clean case. \section{Numerical simulations}\label{app.num} To develop a precise theoretical description of the subgap region, we solve Eqs.\ (2) of the main paper numerically for the two-dimensional and three-dimensional cases. The main element of these equations is the Green's function in real space, $\hat{g}_{\alpha }(r)$. At $r=0$ the Green's function does not depend on dimensionality, \[ \hat{g}_{\alpha}(0)=\frac{-E+\Delta_{\alpha}\hat{\tau}_{x}}{\sqrt {\Delta_{\alpha}^{2}-E^{2}}}. \] Large-$r$ asymptotics of $\hat{g}_{\alpha }(r)$ for $k_{F,\alpha}r\gg1$ are given by \begin{widetext} \begin{align*} \hat{g}_{\alpha }(r)& =\left[ \frac{-E+\Delta _{\alpha }\hat{\tau}_{x}}{ \sqrt{\Delta _{\alpha }^{2}-E^{2}}}\cos \left( k_{F,\alpha }r-\frac{\pi }{4} \right) -\hat{\tau}_{z}\sin \left( k_{F},_{\alpha }r-\frac{\pi }{4}\right) \right] \frac{\exp \left( -\sqrt{\Delta _{\alpha }^{2}-E^{2}}r/v_{F,\alpha }\right) }{\sqrt{\pi k_{F,\alpha }r/2}}\text{, for 2D case} \\ \hat{g}_{\alpha }(r)& =\left[ \frac{-E+\Delta _{\alpha }\hat{\tau}_{x}}{ \sqrt{\Delta _{\alpha }^{2}-E^{2}}}\sin (k_{F,\alpha }r)-\hat{\tau}_{z}\cos (k_{F,\alpha }r)\right] \frac{\exp (-\sqrt{\Delta _{\alpha }^{2}-E^{2}} r/v_{F,\alpha})}{k_{F,\alpha }r}\text{, for 3D case} \end{align*} \end{widetext} As the probability to find two impurities at distance $\sim 1/k_{F}$ is very small, the structure of states is mostly determined by these asymptotics and the value at $r=0$. To match the large-$r$ asymptotics with the $r=0$ value, we use an approximate forms of the Green's functions. For the 2D case we use \begin{align} \hat{g}_{\alpha}(r) & =\left[ \frac{-E+\Delta_{\alpha}\hat{\tau}_{x}} {\sqrt{\Delta_{\alpha}^{2}-E^{2}}}J_{0}\left( k_{F,\alpha}r\right) +\hat{\tau}_{z}J_{1}\left( k_{F,\alpha}r\right) \right] \nonumber \\ & \times\exp\left( -\sqrt{\Delta_{\alpha}^{2}-E^{2}}r/v_{F,\alpha}\right) ,\label{ApproxGreens2D} \end{align} where $J_{0}\left( x\right) $ and $J_{1}\left( x\right) $ are Bessel functions and for the 3D case we use \begin{align} \hat{g}_{\alpha}(r) & \! =\!\left[ \frac{-E\!+\!\Delta_{\alpha}\hat{\tau }_{x}} {\sqrt{\Delta_{\alpha}^{2}\!-\!E^{2}}}\frac{\sin(k_{F,\alpha} r)}{k_{F,\alpha} r}-\hat{\tau}_{z} \mathcal{U}(k_{F,\alpha}r)\frac {\cos(k_{F,\alpha}r)}{k_{F,\alpha}r} \right] \nonumber\\ & \times\exp(-\sqrt{\Delta_{\alpha}^{2}-E^{2}}r/v_{F,\alpha}) \label{ApproxGreens3D} \end{align} with the interpolation function $\mathcal{U}(z)=z^{2}/(z^{2}+1)$. We expect that the exact behavior of the Green's functions at $k_{F,\alpha}r\!\sim\! 1$ not reproduced by these interpolations have little influence on the properties of the subgap states. In this paper we limit ourselves with the simplest case of two equivalent bands, meaning that $k_{F,1}=k_{F,2}\equiv k_{F}$, $v_{F,1}=v_{F,2}\equiv v_{F}$, and $\Delta_1=-\Delta_2\equiv \Delta$. The gap $\Delta$ is used as a unit of energy and $k_F^{-1}$ is used as a unit of length. We employ the following numerical procedure. First, we define an impurity realization by $N_{i}$ random coordinates [$\mathbf{R}_{l}$] in a box, $0<R_{l,a}<R$, and random impurity signs $\delta_{l}=\pm1$ for the off-diagonal scattering amplitudes, $U_{12}^l=\delta_{l}U_{12}$. From the linear $4N_{i}\times4N_{i}$ system defined by Eq.\ (2) of the main paper, we find the eigenenergies, $E_{\lambda}$, and corresponding eigenstates $\Psi^{\lambda}(\mathbf{R}_{l})$. From the set of eigenenergies, $E_{\lambda}$, we compute the average density of state \begin{equation} N_{s}(E)=\left\langle \sum_{\lambda}\delta(E-E_{\lambda})\right\rangle , \end{equation} where average is taken over many impurity realizations. Practically, this implies that for every realization we find the number of states $\Delta N(E)$ falling within the energy interval $[E-\Delta E/2,E+\Delta E/2]$ and then compute the average $N_{s}(E)=\langle\Delta N(E)/(\Delta ER^{2})\rangle$. As an isolated impurity generates one localized state, for small concentration of impurities the normalization condition $\int_{0}^{\Delta}N_{s}(E)dE=N_{i} /R^{2}$ is satisfied. We normalize $N_{s}(E)$ to the total normal density of states for excitations, $N_{n}=4\nu$, where for the 2D case the single-band DoS per electron is given by $\nu=k_{F}/(2\pi v_{F})$ and for the 3D case, $\nu=k_{F}^{2}/(2\pi^{2}v_{F})$. To characterize localization properties, we also compute the average confinement length for states at given energy, \begin{equation} L_{\mathrm{cnf}}(E,R)=\left\langle \sqrt{\sum_{a}\left\langle \delta r_{a} ^{2}\right\rangle _{\lambda}}\right\rangle _{E_{\lambda}=E} , \end{equation} where $a=x,y[,z]$ and the confinement length of state $\lambda$, $\left\langle \delta r_{a}^{2}\right\rangle _{\lambda}$, is determined by its wave function $\Psi_{l}^{\lambda}$ as \begin{align*} \left\langle \delta r_{a}^{2}\right\rangle _{\lambda} & =\left\langle r_{a}^{2}\right\rangle _{\lambda}-\left\langle r_{a}\right\rangle _{\lambda }^{2},\\ \left\langle r_{a}\right\rangle _{\lambda} & =\sum_{l}R_{l,a}|\Psi _{l}^{\lambda}|^{2},\ \ \left\langle r_{a}^{2}\right\rangle _{\lambda} =\sum_{l}R_{l,a}^{2}|\Psi_{l}^{\lambda}|^{2} \end{align*} with $|\Psi_{l}^{\lambda}|^{2}\!=\!\sum_{\alpha}\left[ |u_{\alpha}^{\lambda }(\mathbf{R}_{l})|^{2}\!+\!|v_{\alpha}^{\lambda}(\mathbf{R}_{l})|^{2}\right] $ and $\sum_{l}|\Psi_{l}^{\lambda}|^{2}\!=\!1$. The behavior of the confinement length with increasing system size, $R$, determines wether states at given energy are localized or not. For delocalized states, $L_{\mathrm{cnf}}(E,R)$ is limited by the system size and grows proportionally to $R$. For localized states, $L_{\mathrm{cnf}}(E,R)$ saturates at a finite value, which gives the average localization length of states with energy $E$, $ \lim_{R\rightarrow\infty}L_{\mathrm{cnf}}(E,R)=L_{\mathrm{loc}}(E). $ In three dimensions for relatively small concentration of impurities one can expect the existence of a mobility edge in the subgap region separating localized and delocalized states. It is defined as the energy $E_{\mathrm{ME}}(n_{i}, \gamma_{\alpha\beta})$ at which the localization length diverges, $L_{\mathrm{loc}}(E\! \rightarrow\! E_{\mathrm{ME}})\rightarrow\infty$. At a critical concentration of impurities depending on the scattering parameters, $n_{cr}(\gamma_{\alpha\beta})$, the mobility edge reaches zero energy, $E_{\mathrm{ME}}[n_{cr}(\gamma_{\alpha\beta}), \gamma_{\alpha\beta}]=0$, and all states become delocalized. In the calculations of the concentration dependences presented in the Figs.\ 4a,b we took into account suppression of the average gap parameter described by Eqs.\ (\ref{GapSCTM1}) and (\ref{GapSCTM2}) and corresponding increase of the coherence length $\xi=v_F/\Delta$. \section{Density of states at small concentration of impurities: pairs-dominated regime}\label{app.pairs} \begin{figure}[ptb] \begin{center} \includegraphics[width=3.2in]{DoSNi10Epair.eps} \end{center} \caption{(color online) \emph{Left:} The 2D DoS for small concentration of impurities, same as in Fig.1 of the main paper. \emph{Right:} The energy of impurity pair as a function of the distance between impurities for same-sign and opposite-sign impurities. The arrows mark the extremum points which account for the small peaks in the DoS.} \label{Fig-DoS-pairs} \end{figure} At small concentrations of impurities, $n_{i}\xi^{d}\ll1$, the density of state is determined by impurity pairs.\cite{LifshitzPGBook} The interaction between two close impurities gives the correction to the energy\cite{RusinovPZheTF68}, $E_{\mathrm{pair}}(r)=\varepsilon_0+\delta \varepsilon (r)$ and the behavior of the correction $\delta \varepsilon (r)$ depends on the relative sign of the off-diagonal scattering potential $U_{12}$ for two impurities. For same-sign impurities the single-site energy level splits into two levels corresponding to symmetric and antisymmetric combinations of the wave functions at the impurity sites. The separation-dependent energy corrections, $\delta\varepsilon_{\pm}$, rapidly oscillate with distance between the impurities $r$ as \[ \delta\varepsilon_{\pm}(r)\propto\pm\frac{\sin(k_{F}r+\alpha_{d})} {(k_{F}r)^{(d-1)/2}}\exp\left( -\sqrt{1-\varepsilon_{0}^{2}}r/\xi\right), \] where $\xi=v_{F}/\Delta$ is the coherence length and $d$ is space dimensionality, see Fig.\ \ref{Fig-DoS-pairs}. For opposite-sign impurities the energy level remains double-degenerate with two states corresponding to localization near two impurity sites. The separation-dependent energy shift in this case is always positive and much smaller than for the same-sign impurities, \[ \delta\varepsilon(r)\propto\frac{\sin^{2}(k_{F}r+\alpha_{d})}{(k_{F}R)^{d-1} }\exp\left( -2\sqrt{1-\varepsilon_{0}^{2}}r/\xi\right). \] The coefficients in the energy corrections depend on the scattering parameters of impurities. The contribution to the DoS coming from impurity pairs with separation $r_{p}$ less than the typical distance, $r_{p}\ll n_{i}^{-d}$, can be evaluated in a simple way.\cite{LifshitzPGBook} The concentration of the impurity pairs with separations between $r_{p}$ and $r_{p}+dr_{p}$ is given by $\frac{A_{d}}{2}n_{i}^{2} r_{p}^{d-1}dr_{p}$ with $A_{2}=2\pi$ and $A_{3}=4\pi$. The contribution to the DoS at the energy $E$ is given by the pairs satisfying the equation $E=E_{\mathrm{pair}} (r_p)$ and for a nonmonotonic dependence $E_{\mathrm{pair}}(r)$, this equation may have several solutions. The pair contribution to the DoS can be evaluated as \[ N_{s}(E)=\frac{A_{d}}{2}n_{i}^{2}\sum_{r_p}\left[r_{p}(E)\right] ^{d-1}\left\vert \frac{dr_{p}}{dE}\right\vert , \] where the sum is taken over all values of $r_p$ corresponding to the same energy. For pair separations corresponding to the energy extrema, $E_{\mathrm{pait}}(r)\approx E_{e}+\frac{a} {2}(r-r_{e})^{2}$, the isolated-pairs approximation gives a divergency at $E\rightarrow E_e$, \[ N_{s}(E)\approx\frac{A_{d}n_{i}^{2}\left[ r_{e}\right] ^{d-1}} {\sqrt{2a\left( E-E_{e}\right) }}. \] These singularities, smeared by interactions with more remote impurities, account for the small peaks found in the DoS at small concentrations, see Fig. \ref{Fig-DoS-pairs}. The energy for opposite-sign impurities has a series of minima at exactly $E=\varepsilon_{0}$ meaning that there is also such a pair singularity at the peak center. This explains the sharpness of the peak at small concentrations of impurities.
\section{Buckling Loads} \label{s3} One checks that $(y,M,\theta,Q)=(0,0,0,0)$ is a solution to \eqref{c31}, \eqref{c52} for all loads $\nu$. This solution corresponds to a flat sheet parallel to the substrate at the equilibrium spacing $\bar{d}$. In this section, we find the loads at which the sheet buckles from the flat configuration and we describe how these buckling loads depend on various parameters in the problem. Linearizing \eqref{c31} about the trivial branch yields \begin{subequations} \label{c12} \begin{align} y' &= \theta, \label{c13}\\ M' &= -\nu \theta - Q, \label{c14}\\ \theta' &= M, \label{c15}\\ Q' &= -f_{1}y, \label{c16} \end{align} \end{subequations} where $f_{1} \ensuremath{ := } f'(0)=(L^{4}/\alpha)\bar{f}'(\bar{d})$ by $\subref{c26}{3}$. From \eqref{c39}, \eqref{b9}, we have that $\bar{f}'(\bar{d})=-.12$ nN/nm$^{3}$ for $\bar{f}=\bar{f}_{I}$ and that $\bar{f}'(\bar{d})=-84.1$ nN/nm$^{3}$ for $\bar{f}=\bar{f}_{II}$. The linearized boundary conditions for \eqref{c12} are the same as \eqref{c52}. To locate critical loads, it is convenient though not essential to reformulate \eqref{c12} as a single fourth-order equation. We do so by noting that \begin{equation} \theta'''' = M''' = -\nu \theta'' - Q'' = -\nu\theta'' + f_{1}y' = -\nu\theta'' + f_{1}\theta, \label{c17} \end{equation} and hence \begin{equation} \theta'''' + \nu\theta'' - f_{1}\theta = 0. \label{c18} \end{equation} By \eqref{c15}, the condition \eqref{c50} corresponds to $\theta'(0)=\theta'(1)=0$. Also, by \eqref{c14} and \eqref{c15}, $Q=-\nu\theta-\theta''$, and then by \eqref{c16} $f_{1}y=\nu\theta'+\theta'''$. Hence \eqref{c52} implies that \begin{subequations} \label{c19} \begin{gather} \theta'(0)=\theta'(1)=0, \text{\ and}\label{c53}\\ \theta'''(0) = \theta'''(1) = 0 \quad\text{or}\quad \theta''(0) + \nu \theta(0) = \theta''(1) + \nu \theta(1) = 0 \label{c54} \end{gather} \end{subequations} are the two sets of boundary conditions for \eqref{c18}. The critical loads are the values of $\nu$ at which \eqref{c18}, \eqref{c19} has a non-trivial solution. For a given $\nu$, \eqref{c18}, \eqref{c19} has a non-trivial solution if and only if \eqref{c12}, \eqref{c52} has a non-trivial solution. To find non-trivial solutions of \eqref{c18}, \eqref{c19}, we compute the characteristic roots of \eqref{c18}, which are $\pm \sqrt{(-\nu\pm\sqrt{\nu^{2}+4 f_{1}})/2}$. Using these roots, we find the general solution to \eqref{c18}, to which we apply one set of boundary conditions from \eqref{c19}. This yields a system of linear equations $M(\nu)\mathbf{c}=\mathbf{0}$ for the vector $\mathbf{c}\in\realsTo{4}$ of arbitrary constants in the general solution. Here $M(\nu)$ is a $4\times 4$ matrix whose entries depend nonlinearly on $\nu$. The buckling loads are the values of $\nu$ such that $M(\nu)\mathbf{c}=\mathbf{0}$ has non-trivial solutions. Hence it is sufficient to consider $\nu\mapsto\mydet{M(\nu)}$. Performing the computations just described yields \begin{equation} \mydet{M(\nu)} = \sin\lambda_{1}\sin\lambda_{2} \label{c55} \end{equation} for the boundary conditions \eqref{c53}, $\subref{c54}{1}$ and \begin{equation} \mydet{M(\nu)} = \sqrt{-f_{1}}\lambda_{2} \Biggl( 2(\cos\lambda_{1}\cos\lambda_{2}-1) + \frac{\lambda_{1}^{6}+\lambda_{2}^{6}}{(-f_{1})^{3/2}} \sin\lambda_{1}\sin\lambda_{2} \Biggr) \label{c33} \end{equation} for the boundary conditions \eqref{c53}, $\subref{c54}{2}$, where \begin{equation} \lambda_{1} = \sqrt{(\nu -\sqrt{\nu^{2}+4 f_{1}})/2},\qquad \lambda_{2} = \sqrt{(\nu +\sqrt{\nu^{2}+4 f_{1}})/2}. \label{c34} \end{equation} The expressions \eqref{c55} and \eqref{c33} are correct for $\nu>2\sqrt{-f_{1}}$. One can show that there are no non-trivial solutions for $0<\nu\leq2\sqrt{-f_{1}}$. Note that $f_{1}$ and hence $\mydet{M(\nu)}$ depend on $\alpha$ and $L$. Also, note that \eqref{c55} implies that either $\lambda_{1}=m\pi$ or $\lambda_{2}=m\pi$, which is equivalent to \begin{equation} \nu = \pi^{2}m^{2}-f_{1}/\pi^{2}m^{2}, \label{c56} \end{equation} a formula that could be found more directly by substituting $\theta=A\cos(m\pi s)$ into \eqref{c18}, where $A$ is an arbitrary amplitude. For a given length $L$, we let $\nu^{*}$ denote the buckling load, i.e., the smallest critical load. In the rest of this section, we illustrate how $\nu^{*}$ depends on the interaction force, the boundary conditions, the length of the sheet $L$, and the bending stiffness $\alpha$. These results are computed using either \eqref{c55} or \eqref{c33}. For all the results described below, we take $\alpha=.16$ nN nm unless otherwise noted. Figure~\ref{f8} indicates how $\nu^{*}$ depends on the length $L$ of the sheet for graphene supported by an SiO$_{2}$ substrate. Figure~\ref{f8} is for hinged boundary conditions \eqref{c50}, $\subref{c51}{1}$. From the figure we see that as $L$ increases the buckling load oscillates. The local minima of these oscillations equal $.277$ nN/nm, the local maxima decrease, and $\nu^{*}$ converges to the value of the local minima as $L$ gets large. See also Figure~\ref{f7} (a) below. This oscillatory behavior is well-known. See \cite{gjs&dhh:b_fund}. It occurs because the deformation mode of the buckled solutions changes at the cusps in Figures~\ref{f8} (a), (b). In particular, across the cusps, the nodal behavior and the number of half-sine waves in the deformation mode change (see Figure~\ref{f9}, which is described in the next paragraph). Each deformation mode has a different critical load, and which mode has the lowest critical load changes at the cusps. \begin{figure}[h] \begin{minipage}{1\linewidth} \includegraphics[width=1\linewidth]% {sg_ym_alpha_p16_feval_plot.png} \end{minipage} \caption{Buckling loads for SiO$_{2}$ substrate and hinged boundary conditions. Figures (a) and (b) show how the buckling load, denoted by $\nu^{*}$, depends on $L$ for $\alpha=.16$ nN nm.} \label{f8} \end{figure} Figure~\ref{f9} illustrates how the nodal behavior of the buckled solutions changes at the values of $L$ that correspond to the cusps in Figures~\ref{f8} (a), (b). For the curve describing $\nu^{*}$ as a function of $L$, there are cusps at $L=4.8$, $8.3$, $11.7$, and $15$. The second and third of these cusps can be seen in Figure~\ref{f9} (a). Figure~\ref{f9} (b) shows the buckled configuration corresponding to the buckling load $\nu^{*}$ for $L=6$ nm. The curve in Figure~\ref{f9} (b) has a single simple zero, or node, on $(0,L)$, which is the case for all buckled solutions corresponding to lengths $L$ between $4.8$ and $8.3$. Figures~\ref{f9} (c) and (d) show how the number of nodes of the buckled solutions increases for solutions corresponding to $L$ between $8.3$ and $11.7$ and for solutions corresponding to $L$ between $11.7$ and $15$. \begin{figure}[h] \hspace*{.1\linewidth} \includegraphics[width=.8\linewidth]% {sg_ym_cusps.pdf} \vspace*{-.05\linewidth} \caption{Shape of buckled solutions. Figure (a) shows how the buckling load $\nu^{*}$ depends on $L$ for $\alpha=.16$ nN nm and for graphene supported on an SiO$_{2}$ substrate with hinged boundary conditions. The points marked I, II, and III correspond to the buckling loads for $L=6$, $9$, and $12$ nm. Figures (b), (c), and (d) show how the nodal behavior of the corresponding buckled solutions changes across the two cusps located between the points labeled I, II, and III in (a). For each solution in (b), (c), and (d), the amplitude has been normalized to 1.} \label{f9} \end{figure} Figure~\ref{f10} indicates how $\nu^{*}$ depends on $L$ for graphene supported by an SiO$_{2}$ substrate and loaded by the floating-edge boundary conditions \eqref{c50}, $\subref{c51}{2}$. As in Figure~\ref{f8}, we see that as $L$ increases the buckling load oscillates. In this case, however, both the local minima and local maxima decrease as $L$ increases. Figure~\ref{f10} (a) suggests that $\nu^{*}$ converges to a limiting value as $L$ gets large. See also Figure~\ref{f7} (a) below. The behavior of the buckled solutions across the cusps is similar to that illustrated in Figure~\ref{f9}. \begin{figure}[h] \begin{minipage}{1\linewidth} \includegraphics[width=1\linewidth]% {sg_qm_alpha_p16_feval_plot.png} \end{minipage} \caption{Buckling loads for graphene supported by an SiO$_{2}$ substrate with floating-edge boundary conditions. Figures (a) and (b) show how the buckling load $\nu^{*}$ depends on $L$ for $\alpha=.16$ nN nm.} \label{f10} \end{figure} Figures~\ref{f6} and \ref{f11} illustrate how $\nu^{*}$ depends on $L$ for graphene supported by a HOPG substrate. Figure~\ref{f6} is for hinged boundary conditions and Figure~\ref{f11} is for floating-edge boundary conditions. Although both Figures~\ref{f6} and Figure~\ref{f11} are qualitatively similar to Figures~\ref{f8} and Figure~\ref{f10}, which are the corresponding figures for graphene on SiO$_{2}$, we note that the predicted buckling loads the HOPG substrate are approximately an order of magnitude larger than the buckling loads for graphene supported by an SiO$_{2}$ substrate. For example, in Figure~\ref{f8} (b), $\nu^{*}=0.2771$ nN/nm for $L=40$ nm while in Figure~\ref{f6} (a), $\nu^{*}= 7.335$ nN/nm for $L=40$ nm. See also Table~\ref{f13}. The difference in the size of the buckling loads predicted by the model is a consequence of the difference between the derivatives of the interaction forces at the equilibrium spacings--- $\bar{f}'_{I}(\bar{d}_{I})=-.12$ nN/nm$^{3}$ for SiO$_{2}$ and $\bar{f}'_{II}(\bar{d_{II}})=-84.1$ nN/nm$^{3}$ for HOPG. Figure~\ref{f12} illustrates how $\nu^{*}$ depends on $\bar{f}'(\bar{d})$ for $L=5$ nm. Note that $-\bar{f}'(\bar{d})$ represents the linear stiffness or `spring constant' of the elastic foundation. \begin{figure}[h] \begin{minipage}{1\linewidth} \includegraphics[width=1\linewidth]% {gg_ym_alpha_p16_feval_plot.png} \end{minipage} \caption{Buckling loads for HOPG substrate with hinged boundary conditions. Figures (a) and (b) show how the smallest eigenvalue, denoted by $\nu^{*}$, depends on $L$ for $\alpha=.16$ nN nm.} \label{f6} \end{figure} \begin{figure}[t] \begin{minipage}{1\linewidth} \includegraphics[width=1\linewidth]% {gg_qm_alpha_p16_feval_plot.pdf} \end{minipage} \caption{Buckling loads for HOPG substrate with floating-edge boundary conditions. Figures (a) and (b) show how the smallest eigenvalue, denoted by $\nu^{*}$, depends on $L$ for $\alpha=.16$ nN nm.} \label{f11} \end{figure} \begin{figure}[b] \hspace*{.15\linewidth} \hspace*{.75in} \includegraphics[width=.4\linewidth]% {f1_f_ev.png} \caption{Buckling loads as a function of the derivative of the interaction force $\bar{f}$ at the equilibrium spacing $\bar{d}$. Both hinged and floating-edge boundary conditions are shown. For each case, $L=5$ nm and $\alpha=.16$ nN nm. The spacing on the horizontal axis is logarithmic.} \label{f12} \end{figure} The last figure in this section, Figure~\ref{f7}, shows how the limiting value of $\nu^{*}$ for large $L$ depends on $\alpha$. Figure~\ref{f7} (a) is for graphene supported by SiO$_{2}$ and for $L=150$ nm. As one would expect, the figure indicates that for large $L$ the buckling load $\nu^{*}$ is the same for both hinged and floating-edge boundary conditions. The figure also shows that the buckling load increases as $\alpha$ increases, i.e., as the sheet gets stiffer. Figures~\ref{f7} (b) corresponds to Figures~\ref{f7} (a) but for graphene supported by a HOPG substrate. \begin{figure}[h] \begin{minipage}{1\linewidth} \includegraphics[width=1\linewidth]% {ym_qm_large_L_feval_plot.png} \end{minipage} \caption{Buckling loads as a function of $\alpha$. Figure (a) is for graphene supported by SiO$_{2}$ and Figure (b) is for graphene supported by HOPG. In both cases, $L=150$ nm.} \label{f7} \end{figure} \bigskip \begin{table} \centering \begin{tabular}{lllllllll} \toprule & \multicolumn{4}{c}{SiO$_{2}$} & \multicolumn{4}{c}{HOPG} \\ \cmidrule(r){2-5} \cmidrule(r){6-9} & \multicolumn{2}{c}{Hinged} & \multicolumn{2}{c}{Floating} & \multicolumn{2}{c}{Hinged} & \multicolumn{2}{c}{Floating} \\ \cmidrule(r){2-3} \cmidrule(r){4-5} \cmidrule(r){6-7} \cmidrule(r){8-9} \multicolumn{1}{c}{$L$} & \multicolumn{1}{c}{min} & \multicolumn{1}{c}{max} & \multicolumn{1}{c}{min} & \multicolumn{1}{c}{max} & \multicolumn{1}{c}{min} & \multicolumn{1}{c}{max} & \multicolumn{1}{c}{min} & \multicolumn{1}{c}{max} \\ Near $5$ nm &0.2770&0.3462&0.4010&0.4617&7.3354&7.4008&7.5417&7.5682\\ Near $25$ nm &0.2770&0.2795&0.2848&0.2858&7.3354&7.3379&7.3448&7.3456\\ Near $150$ nm &0.2770&0.2771&0.2774&0.2774&7.3354&7.3355&7.3358&7.3358\\ \bottomrule \end{tabular} \caption{Summary of Buckling Loads. The min (max) value corresponds to the buckling load at the local minimum (maximum) nearest to the length $L$ on the graph of $\nu^{*}$ as a function of $L$. The numbers in the table have units of nN/nm.}\label{f13} \end{table} The results of this section are summarized in Table~\ref{f13} \section{Conclusion} \label{s4} In this paper we developed a nonlinear continuum model of a graphene sheet supported by a flat rigid substrate and loaded on a pair of opposite edges. We modeled the cross-section of the sheet as an elastica. Using techniques from bifurcation theory, we investigated how the buckling of the sheet depends on the boundary conditions, the composition of the substrate, and the length of the sheet. We also presented numerical results that illustrate some of the possible post-buckling behavior of the sheet. An interesting feature of the post-buckling behavior is that the sheet undergoes snap-buckling, which may have implications for the design of nanoscale devices that use graphene. In our model, we assume the substrate is perfectly flat, which of course is not the case. For example, a substrate like SiO$_{2}$ may have undulations, and several recent papers suggest that graphene supported by an SiO$_{2}$ substrate will form ripples to follow these undulations \cite{deshpande:205411,IshigamiM._nl070613a}. Within the context of our continuum modeling, ripples that form in the sheet prior to loading can be described as initial imperfections. A well-developed branch of bifurcation theory addresses how the presence of such imperfections influences buckling and post-buckling behavior. The study of this question follows naturally from the work presented here. In a second variation of the problem modeled above, we can assume that the graphene sheet is deposited on a deformable, rather than rigid, substrate. Hence the cross-section of the substrate could be modeled as a beam with mechanical properties different from the sheet. This problem is motivated by recent experimental work in which graphene sheets are deposited on substrates of various compositions and strain is induced in the graphene by deforming the substrate \cite{ni:115416,Ni2008,pereira:046801}. This work was supported by NSF under grant number DMS-0407361. \section{Equilibrium Equations} \label{s2} To describe the basic geometry of our problem, we let $\{\mathbf{i},\mathbf{j},\mathbf{k}\}$ denote a right-handed orthonormal basis for $\realsTo{3}$. The rigid substrate is parallel to the $\mathbf{i}\mathbf{k}$-plane, and $\mathbf{j}$ points away from the substrate. See Figure~\ref{f3}. We assume the deformation of the sheet is the same in any cross-section defined by a plane perpendicular to $\mathbf{k}$, and hence the configuration of the sheet is determined by the configuration of a typical cross-section. A cross-section is described by a curve $[0,L]\ni s\mapsto \mathbf{r}(s)$ in the $\mathbf{i}\mathbf{j}$-plane. \begin{figure}[h] \hspace*{.0275\linewidth} \begin{minipage}{4in} \includegraphics{elasticaGeometry.pdf} \end{minipage} \caption{Basic geometry of a typical cross-section.} \label{f3} \end{figure} For the geometry depicted in Figure~\ref{f3}, an appropriate theory of nonlinear rods (see \cite{ssa:b_nonl}) delivers the governing equations \begin{align} &\mathbf{n}' + \mathbf{f} = \mathbf{0}, \qquad M' + \mathbf{k}\cdot(\mathbf{r}'\times\mathbf{n}) = 0, \label{c4} \\ &\mathbf{n}(0)\cdot\mathbf{i} = \mathbf{n}(L)\cdot\mathbf{i} = -\nu, \qquad M(0)=M(L)=0, \label{c1} \end{align} which are the linear and angular momentum balances for the sheet supplemented by boundary conditions. In \eqref{c4}, \eqref{c1}, $\mathbf{n}(s)$ is the contact force per unit width on the material point $s$, $\mathbf{f}(s)$ is the force per unit area exerted by the substrate on $s$, and $M(s)$ is the $\mathbf{k}$ component of the contact torque per unit width on $s$. The boundary condition $\subref{c1}{1}$ describes the load applied to opposite edges of the sheet. The parameter $\nu$ is the component of the load parallel to the substrate at the left edge; $\nu>0$ corresponds to compressive loading. The boundary condition $\subref{c1}{2}$ states that the loaded edges are moment free. Additional boundary conditions are given below. To introduce components, we write \begin{equation} \mathbf{r}(s)=x(s)\mathbf{i}+(y(s)+\bar{d})\mathbf{j}, \qquad \mathbf{n}(s)=P(s)\mathbf{i}+Q(s)\mathbf{j}, \label{c2} \end{equation} where the constant $\bar{d}$ is defined later. We assume the cross-section of the layer is inextensible, so that \begin{equation} \mathbf{r}'(s) = x'(s)\mathbf{i}+y'(s)\mathbf{j} = \cos\theta(s)\mathbf{i}+\sin\theta(s)\mathbf{j} \label{c3} \end{equation} for some function $[0,L]\ni s\mapsto \theta(s)$. Hence $L$ is the length of the sheet. We also assume that $M(s)=\alpha\,\theta'\!(s)$ for some constant $\alpha>0$, which is a material parameter describing the resistance of the sheet to bending. Values from the literature on the continuum modeling of graphene and carbon nanotubes suggest that one can choose $\alpha$ in a range between $.13$ and $.2$ nN nm \cite kim&neto:a_grap, tl&etal:a_ener, PhysRevB.5.4951, dr&db&jm:a_ener, sabio:195409, PhysRevB.46.15546, PhysRevB.65.233407, biy&cjb&jb:a_nan }. To make explicit computations below, we take $\alpha=0.16$ nN nm. Note that inextensibility and a linear dependence of the moment on the curvature are the standard assumptions of the elastica theory. The term $\mathbf{f}$ in $\subref{c4}{1}$ describes the interaction force exerted by the substrate on the sheet. The specific form of $f$ depends on the composition of the substrate. One case we study is a substrate composed of SiO$_{2}$. To determine the interaction force in this case, we start with an expression for the attractive energy between graphene and SiO$_{2}$ from \cite{sabio:195409}. The authors report an attractive energy per unit area of \begin{equation} E = -\frac{\hbar v_{F}}{4a}\frac{g_{1}+g_{2}}{\xi^{2}}, \label{c57} \end{equation} where $g_{1}=5.4\cdot 10^{-3}$, $g_{2}=3.5\cdot 10^{-2}$, $v_{F}=10^{6}$ m/s, $a=.142$ nm , $\hbar$ is Planck's constant, and $\xi$ is the distance between the graphene sheet and the substrate. From \eqref{c57}, we derive an attractive force of the form $-c_{A}/\xi^{3}$, where $c_{A}=1.499\cdot 10^{-2}$ nN nm. Now we define $\mathbf{f}(s)=\bar{f}_{I}(y(s)+\bar{d}_{I})\mathbf{j}$, where $\bar{f}_{I}$ has the form \begin{equation} \bar{f}_{I}(\xi) = \frac{c_{R}}{\xi^{11}} - \frac{c_{A}}{\xi^{3}} \label{c39} \end{equation} with $c_{R}=1.499\cdot 10^{-2}$ nN nm$^{9}$. The repulsive term in \eqref{c39} is derived by first assuming a $\xi^{-11}$ dependence, which is consistent with the repulsive part of the Lennard-Jones potential used to describe the interaction between non-bonded atoms (see \eqref{b9} below), and then choosing $c_{R}$ so that the equilibrium spacing between the sheet and the substrate is $\bar{d}_{I}=1$~nm, which is consistent with some experimental data \cite{akg&ksn:a_rise,ElenaStolyarova05292007}. The graph of $\bar{f}_{I}$ is depicted in Figure~\ref{f2}(a). \begin{figure}[h] \includegraphics[width=1\linewidth]{interaction_terms.pdf} \caption{Interaction forces for (a) SiO$_{2}$ and (b) HOPG substrates. $\bar{d}_{I}=1$ nm and $\bar{d}_{II}=.341$ nm are the unique zeros of $\bar{f}_{I}$ and $\bar{f}_{II}$.} \label{f2} \end{figure} We also consider the case in which the substrate is HOPG. The force described by $\mathbf{f}$ in $\subref{c4}{1}$ arises because of van der Waals\ interactions between the carbon atoms of the sheet and the carbon atoms of the substrate. We consider the interactions between the sheet and only the top layer of the atoms on the HOPG substrate. Hence we model the substrate as a second, rigid sheet of graphene. (Because the van der Waals\ force between carbon atoms decays rapidly as the spacing between the atoms increases, including interactions with additional layers of the HOPG substrate does not significantly change the interaction force between the sheet and the substrate.) To define $\mathbf{f}$ appropriately for a continuum model, we assume the atoms are distributed on the substrate and on the sheet with a uniform atomic density $\sigma=38.177\ \mathrm{nm}^{-2}$, a value computed from the geometry of graphene. Also, we assume the substrate is infinite in extent. By computing an appropriate improper integral, we find that the force per unit area exerted by the substrate on the sheet is $\mathbf{f}(s)=\bar{f}_{II}(y(s)+\bar{d}_{II})\mathbf{j}$, where \begin{equation} \bar{f}_{II}(\xi) = 2\pi \sigma^{2} \left( \frac{c_{12}}{\xi^{11}} - \frac{c_{6}}{\xi^{5}} \right) \label{b9} \end{equation} with $c_{12}= 3.859\times 10^{-9} \ \mathrm{nN}\ \mathrm{nm}^{13}$ and $c_{6} = 2.43\times 10^{-6} \ \mathrm{nN}\ \mathrm{nm}^{7}$. (See \cite{jpw&etal:a_comp} for details. The values of $c_{6}$ and $c_{12}$ are from \cite{lag&ral:a_ener}.) The graph of $\bar{f}_{II}$ is depicted in Figure~\ref{f2}(b). Note that $\bar{f}_{II}$ has a unique zero at $\bar{d}_{II}=(c_{12}/c_{6})^{1/6}=0.341$~nm, which is the equilibrium spacing. Below we denote the $\mathbf{j}$ component of the interaction force by just $\bar{f}$ and indicate whether $\bar{f}=\bar{f}_{I}$ or $\bar{f}=\bar{f}_{II}$ only when presenting numerical results. We treat $\bar{d}$ likewise. Recalling $\subref{c2}{1}$, we see that the $\mathbf{j}$ component of $\mathbf{r}$ is measured from $\bar{d}$, so that $y(s)=0$ implies that $\mathbf{f}(s)=\mathbf{0}$. \begin{subequations} \label{c10} We now reformulate \eqref{c4}, \eqref{c1} in components. First we note that $\mathbf{f}\cdot\mathbf{i}=0$, $\subref{c4}{1}$, and $\subref{c2}{2}$ imply that $P'(s)=0$ for all $s$, and hence $P\equiv -\nu$ by $\subref{c1}{1}$. Then \eqref{c4}, \eqref{c2}, and \eqref{c3} yield the system \begin{align} y' &= \sin \theta, \label{c5}\\ M' &= -\nu \sin \theta - Q\cos \theta, \label{c6}\\ \theta' &= \alpha^{-1}M, \label{c7}\\ Q' &= -\bar{f}(y+\bar{d}). \label{c8} \end{align} We study this system with the boundary conditions \begin{equation} M(0)=M(L)=0, \label{c38} \end{equation} which is $\subref{c1}{2}$, supplemented with either \begin{equation} y(0)=y(L)=0 \qquad \text{or} \qquad Q(0)=Q(L)=0. \label{c9} \end{equation} \end{subequations} For conditions \eqref{c38}, $\subref{c9}{1}$, the loaded edges are moment free and kept at a prescribed distance from the substrate. These conditions correspond to standard hinged boundary conditions. For conditions \eqref{c38}, $\subref{c9}{2}$, the edge is moment free but not kept a prescribed distance from the substrate. Rather, the vertical component of the applied load is prescribed to be zero. Below we shall refer to this second set of conditions as the `floating-edge' boundary conditions. (For boundary conditions $\subref{c9}{1}$, the applied load at the edge may have a non-zero vertical component.) To non-dimensionalize, we define \begin{gather} \hat{s}=\frac{s}{L}, \quad \hat{y}(\hat{s})=\frac{y(s)}{L}, \quad \hat{M}(\hat{s})=\frac{LM(s)}{\alpha}, \quad \hat{\theta}(\hat{s})=\theta(s), \label{c25} \\[2mm] \hat{Q}(\hat{s})=\frac{L^{2}}{\alpha}Q(s), \quad \hat{\nu}=\frac{L^{2}}{\alpha}\nu, \quad \hat{f}(\xi)=\frac{L^{3}}{\alpha}\bar{f}(L\xi+\bar{d}). \label{c26} \end{gather} Upon inserting these rescalings into \eqref{c10} (and dropping the hats), we get \begin{subequations} \label{c31} \begin{align} y' &= \sin \theta, \label{c27}\\ M' &= -\nu \sin \theta - Q\cos \theta, \label{c28}\\ \theta' &= M, \label{c29}\\ Q' &= -f(y), \label{c30} \end{align} \end{subequations} with boundary conditions \begin{subequations}\label{c52} \begin{gather} M(0)=M(1)=0,\quad\text{and}\label{c50}\\ y(0)=y(1)=0 \quad\text{\ or}\quad Q(0)=Q(1)=0.\label{c51} \end{gather} \end{subequations} Note that the problem we have formulated is a version of the classical problem of a beam on an elastic foundation \cite{spt&jmg:b_theo}. Our version differs from most treatments because we do not use the Euler-Bernoulli beam equation. Also, our choice for $\bar{f}$, which describes the interaction of the beam with the foundation, is specific to graphene supported by either an SiO$_{2}$ or HOPG substrate. \section{Introduction} \label{s1} A graphene sheet is a planar hexagonal lattice of carbon atoms with each atom bonded to its three nearest neighbors. Theoretical predictions about the properties of graphene as well as the role graphene plays as the basic structure in other important materials, in particular bulk graphite and carbon nanotubes, have driven efforts to produce isolated single-layer graphene sheets. Somewhat surprisingly, these efforts succeeded only recently, with the discovery of mechanical and chemical methods for isolating individual graphene sheets and functionalized sheets from bulk graphite \cite{ksn&etal:a_elec,ksn&etal:a_twod,hcs&etal:a_func}. The last several years have brought a tremendous amount of theoretical and experimental research on graphene. Much of this recent research explores the novel electronic transport properties of graphene. To a lesser extent, the mechanical properties of graphene have also attracted attention. Transport properties suggest engineering nanoscale devices that use graphene as basic components like nanoscale resonators, switches, and valves. See, for example, \cite{jsb&etal:a_elec}. Mechanical properties suggest the use of graphene in composite materials \cite{rafiee:223103,ss&dd&et_a:grap}. For the latter, understanding the response of individual graphene sheets to applied loads is clearly important. Additionally, for designing graphene-based devices, and more generally for exploiting the transport properties of graphene, an understanding of the mechanical response of graphene may prove essential for at least two reasons. First, to assemble nanoscale components, it may be useful to develop techniques for manipulating individual graphene sheets and related nanoscale structures, which entails understanding how these sheets respond to applied loads \cite{0957-4484-10-3-308,hvr&etal:a_mani,Zhou20073237}. Second, research suggests that the electronic transport properties of graphene are coupled to its mechanical deformation \cite{deshpande:205411,gazit:113411,kim&neto:a_grap,knox:201408,Ni2008,sabio:195409}. In this paper we study the mechanical response of a graphene sheet parallel to and supported by a flat rigid substrate. The sheet is loaded compressively on a pair of opposite edges. See Figure~\ref{f3} below. The problem we formulate can be loosely motivated by a recent experimental paper of Schniepp et al.\@ \cite{hcs&etal:a_bend}, in which functionalized graphene sheets supported on a substrate of highly oriented pyrolytic graphite (HOPG) are manipulated with the tip of an atomic force microscope (AFM). The authors show that the lateral force exerted on the edge of the sheet by the AFM tip can slide the sheet across the substrate or, more interestingly, can fold the edge of the sheet. See Figures~3 and 4 in \cite{hcs&etal:a_bend}. In fact, the sheet can be folded and unfolded by the tip multiple times with this repeated folding occurring along the same location. In our idealized continuum model, the compressive load on the edges of the sheet could describe the lateral load applied by an AFM tip. More generally, the geometry of our problem seems fundamental for the study of the mechanics of graphene because mechanical exfoliation, a technique for isolating individual graphene from bulk graphite, yields graphene sheets supported by a rigid substrate \cite{ksn&etal:a_twod}. Also, several proposed electronic devices are based on graphene in a geometry similar to that of our problem \cite{sabio:195409}. We consider two types of substrates. In one case, the sheet interacts with SiO$_{2}$, a material commonly used to isolate single-layer graphene by mechanical exfoliation \cite{blake:063124}. The interaction between the flexible sheet and the substrate is by van der Waals forces, which act over a short-range between the individual carbon atoms on the sheet and the atoms in the substrate. Several recent papers suggest that the mechanical response of graphene supported on SiO$_{2}$ is significantly influenced by the interaction between the sheet and the substrate \cite{sabio:195409}. As a second case, we consider a graphene sheet interacting by van der Waals forces with a second, rigid graphene sheet. The interaction between two graphene sheets should approximate well the interaction between graphene and the HOPG substrate in the experiments in \cite{hcs&etal:a_bend} (although we note that the specific interaction terms we present in the next section are appropriate for describing a substrate supporting pure graphene, not functionalized graphene). Modeling both the graphene sheet and the substrate as continua, we derive equilibrium equations for the static response of the compressively loaded sheet. We assume that the sheet deforms the same in each cross-section, a simplification supported by the atomistic simulations in \cite{ql&rh:a_nonl,jpw&etal:a_comp}. A typical cross-section of the sheet is modeled as an elastic rod. Our assumptions along with appropriate boundary conditions---discussed in the next paragraph--- yield a geometrically exact version of the classical problem of a beam on a nonlinearly elastic foundation. Two sets of boundary conditions are considered. For the first set, the horizontal component of the applied load at opposite edges of the sheet is prescribed, the loaded edges are moment free, and the loaded edges are kept at a prescribed distance from the substrate. These correspond to classical `hinged' boundary conditions. For the second set of boundary conditions, we again prescribe the horizontal component of the applied load at opposite edges and we require that the loaded edges are moment free. But the loaded edges are not kept a prescribed distance from the substrate. Rather, the vertical component of the applied load is prescribed to be zero. This second set of conditions is loosely motivated by the AFM tip experiments described above, in which the edges of the sheet folded over as the lateral load was applied. Using standard ideas from bifurcation theory, we investigate how the mechanical response of the graphene sheet depends on the length of the sheet, on the boundary conditions, and on the composition of the substrate. In our bifurcation analysis, the trivial branch corresponds to a flat sheet parallel to and at the equilibrium spacing from the substrate. By linearizing about the trivial branch, we compute the critical load at which the sheet buckles. How this critical load depends on length, boundary conditions, and the composition of the substrate is summarized in Table~\ref{f13}. We show that the critical load converges to a nonzero value as the length of the sheet goes to infinity. To illuminate the post-buckling behavior, we numerically continue the branching solutions. Our numerical results indicate that the post-buckling behavior includes secondary bifurcations, and, more interestingly, snap-buckling. There is an extensive literature on the buckling of single-walled and multi-walled carbon nanotubes under axial and radial loads. For results that use continuum modeling, see, for example, \cite{Rdec00,Ru-2001-Mech-Solids,post:a_h-ss}. The literature on buckling of graphene sheets under compressive loading is less extensive. In \cite{jpw&etal:a_comp}, we studied the buckling under edge loads of two parallel, deformable graphene sheets interacting by van der Waals forces. We showed that a simple continuum model of buckling yields predictions that are qualitatively similar to the predictions of atomistic simulations. In \cite{ql&rh:a_nonl}, the authors studied the buckling of graphene nanoribbons under compressive load. The atomistic simulations they performed were based on energy minimization techniques in which the carbon-carbon interactions were modeled by the REBO potential. Their results suggest that linear beam theory predicts reasonably well the buckling of graphene. In their problem, the sheet did not interact with a substrate. In \cite{SakhaeePour2009266}, the author used the `molecular structural mechanics' approach developed in \cite{Li20032487} to perform atomistic simulations of the buckling of a graphene sheet under compressive loading. How the boundary conditions, the length of the sheet, and the aspect ratio of the sheet influenced the buckling load were studied. In their model, the sheet is freely suspended and not interacting with a substrate. A comparison between Figures~2 and 3 and between Figures~4 and 5 in \cite{SakhaeePour2009266} suggests that for sheets with lateral dimensions larger than 25 nm, chirality does not strongly influence the buckling load. That the buckling load does not appear to depend on chirality was also noted in \cite{ql&rh:a_nonl}. In our continuum model, we cannot describe the chirality of the sheet. In \cite{Pradhan2009}, the authors apply nonlocal elasticity theory to study the buckling of graphene sheets under biaxial compression. They study a single graphene sheet that is freely suspended and not interacting with a substrate. Figure~5 in \cite{Pradhan2009} suggests that as the lateral dimensions of the sheet exceed about 25 nm nonlocal effects become less important for determining buckling loads. This observation supports the validity of our results, where nonlocal effects are not considered. We note also the recent paper \cite{Li2010}, in which the authors study the morphology of a graphene sheet supported by a patterned substrate. Their results indicate that as a certain parameter controlling the pattern on the substrate is varied, the sheet can exhibit a `snap-through' instability, which appears similar to the snap-buckling we describe below. An interesting issue related to the basic morphology of a graphene sheet supported on a substrate is the existence of ripples \cite{Meyer2007}. Several theories have been proposed to explain this rippling \cite{deshpande:205411,Fasolino2007,geringer:076102,IshigamiM._nl070613a,ref0295-5075-85-4-46002}. Here we do not directly address the question of rippling. In the next section, we derive the boundary-value-problem for a graphene sheet interacting with a substrate. In Section~\ref{s3}, we identify a trivial branch and compute buckling loads. Section~\ref{s7} contains numerical results that describe post-buckling behavior. The final section summarizes our results and mentions several additional problems that could be studied within the framework of the model we develop. \section{Post-Buckling Behavior} \label{s7} In this section we describe the post-buckling behavior predicted by our model. We present numerical results computed using the bifurcation software AUTO \cite{Auto}. Figure~\ref{f14} depicts a part of the bifurcation diagram for a sheet of length $L=30.4$ nm supported by a HOPG substrate with hinged boundary conditions. (Note that the dashed lines are used for clarity and not to indicate that the branches are unstable.) The variable on the vertical axis, which measures the size of solutions, is the angle $\theta(0)$ between the $\mathbf{i}$-axis and the tangent line at the left end of the sheet. The diagram shows a portion of the trivial branch containing the buckling load $\nu^{*}$ and the next two critical loads. There is a pitchfork bifurcation at $\nu^{*}$. The next two critical loads correspond to transcritical bifurcations. We also observe secondary bifurcations on the upper and lower branches emanating from the third critical load. Figure~\ref{f14} should be compared to Figure~11(b) in \cite{ql&rh:a_nonl} to illustrate the effect of the supporting substrate on the mechanical response of the graphene sheet. \begin{figure}[h] \hspace*{.1\linewidth} \begin{minipage}{1\linewidth} \includegraphics[width=.8\linewidth]% {gg_ym_L_30p4_alpha_p16_bd.png} \end{minipage} \caption{Bifurcation diagram for graphene supported by a HOPG substrate with hinged boundary conditions. $L=30.4$ nm. The first bifurcation is a pitchfork. The dashed lines are used for clarity and not to indicate unstable branches.} \label{f14} \end{figure} In Figure~\ref{f15}, we illustrate a possible post-buckling path based upon the bifurcation diagram in Figure~\ref{f14}. We assume the sheet undergoes a quasi-static loading process in which a compressive edge load is slowly increased from the unloaded configuration. For loads below the buckling load, the trivial branch is stable and the sheet remains flat. We assume that at the buckling load the trivial branch loses stability. Because the pitchfork opens to the left, if the load is further increased the solution must jump to a different branch. For macroscopic structures, this type of jump is often referred to as snap-buckling. Determining rigorously to which branch the solution jumps entails analyzing the stability of branches by considering an appropriate potential energy for \eqref{c4}, \eqref{c1} or by studying the dynamical version of these governing equations. Here we do not undertake this analysis. (See \cite{sdr:a_stab} for a stability analysis based on minimizing potential energy in a related problem.) However, to illustrate one possibility, we show the solution jumping to the lower branch of the upper pitchfork in Figure~\ref{f14}. The solution follows this branch until the secondary bifurcation point on the upper branch emanating from the third critical load in Figure~\ref{f14}. If the load is further increased, another snap-buckling may occur. The shapes of buckled solutions are depicted on the right in Figure~\ref{f15}. Note that both $x$ and $y$ are rescaled to be dimensionless. \begin{figure}[h] \begin{minipage}{1\linewidth} \includegraphics[width=1.\linewidth]% {gg_ym_L_30p4_alpha_p16_snap3.png} \end{minipage} \caption{Possible buckling path. On the left is a part of the bifurcation diagram shown in Figure~\ref{f14}. One possible post-buckling path is indicated. On the right are the shapes of the buckled solutions along this path. The variables $x$ and $y$ are rescaled to be dimensionless.} \label{f15} \end{figure} \newpage Figure~\ref{f16} depicts a part of the bifurcation diagram for a sheet of length $L=30.8$ nm supported by a HOPG substrate with hinged boundary conditions. (Here also the dashed lines are used for clarity and not to indicate that the branches are unstable.) The diagram shows a portion of the trivial branch containing the buckling load $\nu^{*}$ and the next two critical loads. The bifurcation at $\nu^{*}$ is transcritical. The next two critical loads correspond to pitchfork bifurcations. There are secondary bifurcations on the upper and lower branches emanating from the second critical load. In Figure~\ref{f17}, we illustrate a possible post-buckling path based upon the bifurcation diagram in Figure~\ref{f16}. For loads below the buckling load, the trivial branch is stable and the sheet remains flat. At the buckling load the trivial branch loses stability. The solution follows the lower branch from the transcritical bifurcation until a turning point is reached. If the load is further increased the sheet undergoes a snap-buckling and the solution jumps to a different branch. For example, the solution could jump to the lower branch of the upper pitchfork in Figure~\ref{f16}. The solution follows this branch until the secondary bifurcation point on the upper branch emanating from the second critical load. If the load is further increased, another snap-buckling may occur. The shapes of buckled solutions are depicted on the right. Both $x$ and $y$ are rescaled to be dimensionless. \begin{figure}[h] \hspace*{.1\linewidth} \begin{minipage}{1\linewidth} \includegraphics[width=.8\linewidth]% {gg_ym_L_30p8_alpha_p16_bd.png} \end{minipage} \caption{Bifurcation diagram for graphene supported by a HOPG substrate with hinged boundary conditions. $L=30.8$ nm. The first bifurcation is transcritical. The dashed lines are used for clarity and not to indicate unstable branches.} \label{f16} \end{figure} \newpage \begin{figure}[h] \begin{minipage}{1\linewidth} \includegraphics[width=1.\linewidth]% {gg_ym_L_30p8_alpha_p16_bd_post_buck.png} \end{minipage} \caption{Possible post-buckling path. On the left is a part of the bifurcation diagram shown in Figure~\ref{f16}. One possible post-buckling path is indicated. On the right are the shapes of the buckled solutions along this path. The variables $x$ and $y$ are rescaled to be dimensionless.} \label{f17} \end{figure} The main point of the numerical results in this section is to illustrate the possibility of snap-buckling. This possibility suggests that the sheet, if loaded beyond the buckling load, would undergo a rapid and relatively large deformation from the flat configuration. As noted in the introduction, recent experimental and theoretical work on graphene establishes a connection between the deformation of the sheet and its transport properties. Hence our modeling predicts that at the buckling load the sheet could undergo a large, rapid change in transport properties. This property could be exploited in the design of nanoscale devices that incorporate graphene sheets.
\section{Introduction} In cosmology, the sources of matter that drive the expansion of the universe are often considered to be either fluids or, in particular, while considering the inflationary epoch or late time acceleration, as scalar fields. Almost always, the fluids are assumed to be perfect, i.e. it is assumed that they do not possess any intrinsic, non-adiabatic pressure perturbation. In contrast, generically, the non-adiabatic pressure perturbation proves to be non-zero for scalar fields. It is well known that iso-curvature (i.e. the extrinsic, non-adiabatic pressure) perturbations are always present when one considers multiple fields and/or fluids. In the context of inflation, it is often said that single scalar field models are adiabatic (see any of the following texts~\cite{texts} or reviews~\cite{reviews}). While this is true, it is not because the non-adiabatic pressure perturbation is identically zero for the single field models, but because they decay exponentially for cosmological modes after they leave the Hubble radius during inflation (for a discussion on specific cases, see Refs.~\cite{bep-shs}; for a generic discussion, see Ref.~\cite{christopherson-2009}). It is interesting to enquire whether there exist scalar field models wherein the intrinsic, non-adiabatic pressure perturbation vanishes exactly (say, at the first order in the perturbations), as in the case of perfect fluids. In this paper, using the standard definition of the non-adiabatic pressure perturbation associated with the scalar fields, we shall obtain a condition on the Lagrangian density describing the scalar field by demanding that the non-adiabatic pressure perturbation is identically zero at the first order in the perturbation theory. We shall also discuss specific examples of scalar field models that satisfy this condition. Following the convention with fluids, we shall call these fields as perfect scalar fields. It is well known that models that depend only on the kinetic energy of the scalar field behave as perfect fluids. We are also able to construct models which seem to depend on the scalar field, and possess no non-adiabatic pressure perturbation. However, we find that all such models that we consider reduce to pure kinetic models after a suitable redefinition of the field. Interestingly, we shall show that, if the perfect scalar fields are used to drive inflation, they can not lead to features in the inflationary, scalar perturbation spectrum. This paper is organized as follows. In the following section, we shall rapidly outline essential, {\it linear},\/ cosmological perturbation theory. We shall highlight the key equations, and point out the standard definition of the intrinsic, non-adiabatic pressure perturbation. In Sec.~\ref{sec:condition}, by demanding that the non-adiabatic pressure perturbation associated with the scalar field vanishes exactly, we shall obtain a condition on the Lagrangian density of the scalar field. We shall discuss a few specific examples of Lagrangian densities that satisfy this condition in Sec.~\ref{sec:examples}. In Sec.~\ref{sec:ii}, we shall show that, if scalar fields with a vanishing non-adiabatic pressure perturbation are responsible for inflation then, in such cases, the first slow roll parameter is always a monotonically decreasing function of time. We shall argue that this behavior points to the fact that such inflatons will not lead to features in the primordial scalar perturbation spectrum. Finally, we shall close with a few concluding remarks in Sec.~\ref{sec:cr}. A few words on our conventions and notations are in order before we proceed. We shall set $c=1$, but shall display $G$ explicitly. We shall work in a spatially flat, $(3 + 1)$-dimensional Friedmann background, and we shall adopt the metric signature of $(+,-,-,-)$. Note that, while the Greek indices $\mu$ or $\nu$ shall denote the spacetime coordinates, the Latin indices $i$ or $j$ shall denote the spatial coordinates. The sub-script $k$ shall refer to the Fourier component of the perturbations. We shall express all the quantities in terms of either the cosmic or the conformal time coordinates. While an overdot shall denote differentiation with respect to the cosmic time, an overprime shall denote differentiation with respect to the conformal time. It is handy to note that, for any given function, say, $f$, ${\dot f}=(f'/a)$ and ${\ddot f}=\left[\left(f''/a^2\right)-\left(f'\, a'/a^{3}\right)\right]$, where $a$ is the scale factor associated with the Friedmann metric. \section{Scalar perturbations in a Friedmann universe---Key equations and definitions}\label{sec:cpt} In this section, using the equations governing the evolution of the scalar perturbations induced by an arbitrary matter field and the standard definition of the non-adiabatic pressure perturbation, we shall arrive at an expression for the non-adiabatic pressure perturbation associated with the scalar field in terms of the adiabatic and the effective speeds of sound associated with the perturbations. In $(3 + 1)$-dimensions, when no perturbations are present, the spatially flat Friedmann universe is described by the line element \begin{equation} {\rm d} s^2 = {\rm d} t^2-a^{2}(t)\; {\rm d}{\bf x}^2 = a^{2}(\eta)\, \left({\rm d}\eta^{2} - {\rm d}{\bf x}^2\right),\label{eq:f-le} \end{equation} where $t$ is the cosmic time, $a(t)$ is the scale factor, and $\eta=\int [{\rm d} t/a(t)]$ denotes the conformal time. If $\rho$ and $p$ denote the energy density and the pressure of the smooth component of the matter field that is driving the expansion, then the Einstein's equations for the above line-element lead to the following Friedmann equations for the scale factor~$a(t)$: \begin{equation} H^2= \left(\frac{8\,\pi\, G}{3}\right)\, {\rho}\qquad{\rm and}\qquad \left(\frac{\ddot a}{a}\right) = -\left(\frac{4\,\pi\, G}{3}\right)\, \left({\rho}+3\, p\right), \end{equation} where $H=({\dot a}/a)$ is the Hubble parameter. \subsection{Equations governing the scalar perturbations and the definition of the non-adiabatic pressure perturbation} If we now take into account the scalar perturbations to the background metric~(\ref{eq:f-le}), then the Friedmann line-element, in general, can be written as~\cite{texts,reviews} \begin{equation} {\rm d} s^2 = (1+2\, A)\,{\rm d} t ^2 - 2\, a(t)\, (\partial_{i} B )\; {\rm d} t\; {\rm d} x^i\, -a^{2}(t)\; \left[(1-2\, \psi)\; \delta _{ij}+ 2\, \left(\partial_{i}\, \partial_{j}E \right)\, {\rm d} x^i\, {\rm d} x^j\right],\label{eq:f-le-sp} \end{equation} where $A$, $B$, $\psi$ and $E$ are the scalar functions that describe the perturbations. The gauge-invariant Bardeen variables that characterize the two degrees of freedom describing the scalar perturbations are given by~\cite{bardeen-1980} \begin{equation} \Phi \equiv A + \left[a\, (B-a\, {\dot E})\right]^{\cdot} \qquad {\rm and}\qquad \Psi \equiv \psi - \left[a\, H\, (B-a\, {\dot E})\right].\label{eq:bps} \end{equation} In the absence of anisotropic stresses, as it is in the case of the scalar field sources that we shall be interested in, it can be readily shown that, at the linear order in the perturbations, the non-diagonal, spatial components of the Einstein's equations lead to the relation: $\Phi=\Psi$. The remaining first order Einstein's equations then reduce to~\cite{texts,reviews} \begin{eqnarray} \left(\frac{1}{a^{2}}\right)\, \nabla^2 \Phi -3\, H\, \left({\dot \Phi} +H\, \Phi\right) &=& \left(4\, \pi\, G\right)\, \left[\delta \rho+ ({\dot \rho}\, a)\; (B-a\, {\dot E})\right] = \left(4\, \pi\, G\right)\, \widehat{\delta \rho},\label{eq:fo-ee-00}\\ \partial_{i}\left({\dot \Phi}+ H\, \Phi\right) &=& \left(4\,\pi\, G\right)\, \partial_{i}\left[\delta \sigma+ [(\rho+p)\, a]\, (B-a\, {\dot E})\right] = \left(4\, \pi\, G\right)\, \left(\partial_{i}\;\widehat{\delta \sigma}\right),\;\;\label{eq:fo-ee-0i}\\ {\ddot \Phi}+ 4\, H\, {\dot \Phi} +\left(2\, {\dot H}+ 3\, H^2\right)\,\Phi\; &=& \left(4\, \pi\, G\right)\, \left[\delta p+ ({\dot p}\, a)\; (B-a\, {\dot E})\right] =\left(4\, \pi\, G\right)\, \widehat{\delta p},\label{eq:fo-ee-ii} \end{eqnarray} where $\delta\rho$, $\delta \sigma$ and $\delta p$ denote the perturbations at the linear order in the energy density, flux, and the pressure of the matter field, respectively, while the quantities $\widehat{\delta \rho}$, $\widehat{\delta \sigma}$ and $\widehat{\delta p}$ represent the corresponding gauge-invariant quantities. The first and the third of the above first order Einstein equations can be combined to lead to the following differential equation for the Bardeen potential $\Phi$~\cite{texts,reviews}: \begin{equation} \Phi^{\prime\prime} +3\, {\mathcal H}\, \left(1+c_{_{\rm A}}^2\right)\, \Phi^{\prime} -c_{_{\rm A}}^2\, \nabla^{2}\Phi + \left[2\, {\mathcal H}'+ \left(1+3\, c_{_{\rm A}}^2\right)\, {\mathcal H}^2\right]\, \Phi = \left(4\, \pi\, G\,\, a^2\right)\, \delta p^{_{\rm NA}},\label{eq:em-bp} \end{equation} where ${\cal H}=(H\, a)$ is the conformal Hubble parameter. In arriving at this equation, we have changed over to the conformal time coordinate, and have made use of the following standard definition of the non-adiabatic pressure perturbation~$\delta p^{_{\rm NA}}$~\cite{gordon-2001}: \begin{equation} \delta p^{_{\rm NA}}=\left(\widehat{\delta p}- c_{_{\rm A}}^2\, \widehat{\delta\rho}\right) = \left(\delta p- c_{_{\rm A}}^2\, \delta\rho\right),\label{eq:delta-p-na-gen} \end{equation} where $c_{_{\rm A}}\equiv\sqrt{\left(p'/\rho'\right)}$ denotes the adiabatic speed of the perturbations. \subsection{Perturbations induced by a generic scalar field} Consider a generic scalar field~$\phi$ that is described by the action~\cite{ki} \begin{equation} S[\phi]=\int\!{\rm d}^{4}x\, \sqrt{-g}\; {\cal L}(X,\phi), \end{equation} where $X$ is a term that describes the kinetic energy of the scalar field, and is defined as \begin{equation} X=\left(\frac{1}{2}\right)\, \left(\partial_{\mu}\phi\; \partial^{\mu}\phi\right). \end{equation} Let us assume that the Lagrangian density ${\cal L}$ is an arbitrary function of the kinetic term $X$ and the field $\phi$. The stress-energy tensor associated with the above action can be written as \begin{equation} T^{\mu}_{\nu} =\left(\frac{\partial{\cal L}}{\partial X}\right)\, \left(\partial^{\mu}\phi\; \partial_{\nu}\phi\right) - \delta^{\mu}_{\nu}\, {\cal L}.\label{eq:set} \end{equation} When no perturbations are present, the energy density and the pressure associated with the homogeneous scalar field are given by \begin{equation} \rho = \left(\frac{\partial {\cal L}}{\partial X}\right)\, (2\, X)- {\cal L} \qquad{\rm and}\qquad p = {\cal L},\label{eq:rho-p-gsf} \end{equation} with $X=({\dot \phi}^{2}/2)$. If we now denote the perturbation in the scalar field as $\delta \phi$, then, the perturbations in the energy density, the momentum flux and the pressure of the scalar field can be obtained to be~\cite{ki,sanil-2008} \begin{eqnarray} \delta\rho &=& \left[\left(\frac{\partial {\cal L}}{\partial X}\right) + \left(2\, X\right)\, \left(\frac{\partial^{2} {\cal L}}{\partial X^{2}}\right)\right]\, {\dot \phi}\,\left(\dot{\delta\phi} - {\dot \phi}\,A\right) - \left[\left(\frac{\partial {\cal L}}{\partial \phi}\right) - \left(2\, X\right)\, \left(\frac{\partial^{2} {\cal L}}{\partial X\, \partial\phi}\right)\right]\,\delta\phi, \label{eq:delta-rho-gsf}\\ \delta\sigma &=& \left(\frac{\partial {\cal L}}{\partial X}\right)\, \left({\dot \phi}\; \delta \phi\right), \label{eq:delta-sigma-gsf}\\ \delta p &=& \left(\frac{\partial {\cal L}}{\partial X}\right)\, {\dot \phi}\,\left(\dot{\delta\phi} - {\dot \phi}\, A\right) +\left(\frac{\partial {\cal L}}{\partial \phi}\right)\, \delta\phi. \label{eq:delta-p-gsf} \end{eqnarray} The gauge-invariant perturbation in the scalar field, say, $\delta \varphi$, is given by~\cite{texts,reviews} \begin{equation} \delta \varphi = \left[\delta \phi+ ({\dot \phi}\; a)\; (B-a\, {\dot E})\right]. \end{equation} The gauge-invariant perturbations in the energy density, the momentum flux and the pressure of the scalar field can be expressed in terms of the gauge-invariant perturbation in the scalar field~$\delta \varphi$ and the Bardeen potential~$\Phi$ as follows: \begin{eqnarray} \widehat{\delta\rho} &=& \left[\left(\frac{\partial {\cal L}}{\partial X}\right) + \left(2\, X\right)\,\left(\frac{\partial^{2} {\cal L}}{\partial X^{2}}\right)\right]\, {\dot \phi}\,\left(\dot{\delta{\varphi}} - {\dot \phi}\, \Phi\right)\, - \left[\left(\frac{\partial {\cal L}}{\partial \phi}\right) - \left(2\, X\right)\,\left(\frac{\partial^{2} {\cal L}}{\partial X\, \partial\phi}\right)\right]\, \delta{\varphi}, \label{eq:gi-delta-rho-gsf}\\ \widehat{\delta\sigma} &=& \left(\frac{\partial {\cal L}}{\partial X}\right)\, \left({\dot \phi}\; \delta \varphi\right), \label{eq:gi-delta-sigma-gsf}\\ \widehat{\delta p} &=& \left(\frac{\partial {\cal L}}{\partial X}\right)\, {\dot \phi}\,\left(\dot{\delta\varphi} - {\dot \phi}\,\Phi\right) +\left(\frac{\partial {\cal L}}{\partial \phi}\right)\, \delta\varphi. \label{eq:gi-delta-p-gsf} \end{eqnarray} Our goal now is to arrive at an expression for the non-adiabatic pressure perturbation $\delta p^{_{\rm NA}}$ for the scalar field in terms of the Bardeen potential~$\Phi$. Algebraically, we find that, one useful way would be to derive the equivalent of the Bardeen equation~(\ref{eq:em-bp}) for the case of the scalar field. On substituting the above expressions for~$\widehat{\delta \rho}$ and $\widehat{\delta p}$ in the first order Einstein's equations~(\ref{eq:fo-ee-00}) and ({\ref{eq:fo-ee-ii}), we find that the Bardeen potential $\Phi$ induced by the perturbations in the scalar field satisfies the following differential equation: \begin{equation} \Phi^{\prime\prime} +3\, {\mathcal H}\, \left(1+c_{_{\rm A}}^2\right)\, \Phi^{\prime} -c_{_{\rm A}}^2\, \nabla^{2}\Phi + \left[2\, {\mathcal H}'+ \left(1+3\, c_{_{\rm A}}^2\right)\, {\mathcal H}^2\right]\, \Phi= \left(c_{_{\rm S}}^2-c_{_{\rm A}}^2\right)\, \nabla^{2}\Phi,\label{eq:em-bp-gsf} \end{equation} where the quantity $c_{_{\rm S}}$ is often referred to as the effective speed of the perturbations, and is given by~\cite{ki} \begin{equation} c_{_{\rm S}}^{2} = \left[\frac{\left({\partial {\cal L}}/{\partial X}\right)}{\left({\partial {\cal L}}/{\partial X}\right) + \left(2\, X\right)\, \left({\partial^{2} {\cal L}}/{\partial X^{2}}\right)}\right]. \label{eq:cs2} \end{equation} It should be mentioned that, in arriving at the above equation for the Bardeen potential~$\Phi$, we have made use of the following equation of motion for the background scalar field~$\phi$: \begin{equation} \left[\left(\frac{\partial {\cal L}}{\partial X}\right) + \left(2\, X\right)\, \left(\frac{\partial^{2}{\cal L}}{\partial X^{2}}\right)\right]\, {\ddot \phi}+\left[\left(3\, H\right)\, \left(\frac{\partial {\cal L}}{\partial X}\right) + {\dot \phi}\, \left(\frac{\partial^{2}{\cal L}}{\partial X\, \partial\phi}\right)\right]\, {\dot \phi} -\left(\frac{\partial {\cal L}}{\partial \phi}\right)=0,\label{eq:em-gsf} \end{equation} which, in turn, can be derived from the equation describing the conservation of the background energy density, viz. \begin{equation} {\dot \rho}+\left(3\, H\right)\; \left(\rho+p\right)=0,\label{eq:ce} \end{equation} and using the expressions for~$\rho$ and~$p$ from Eq.~(\ref{eq:rho-p-gsf}). Upon comparing the equations~(\ref{eq:em-bp}) and~(\ref{eq:em-bp-gsf}), it is straightforward to see that the non-adiabatic pressure perturbation $\delta p^{_{\rm NA}}$ for the scalar field can be expressed as \begin{equation} \delta p^{_{\rm NA}} = \left(\frac{c_{_{\rm S}}^{2} - c_{_{\rm A}}^{2}}{4\, \pi\, G\, a^{2}}\right)\; \nabla^{2}\Phi.\label{eq:delta-p-na-gsf} \end{equation} This result for $\delta p^{_{\rm NA}}$ can be obtained in a more straightforward manner by first noticing that the quantities~$\widehat{\delta \rho}$, $\widehat{\delta \sigma}$ and~$\widehat{\delta p}$ as given by the expressions~(\ref{eq:gi-delta-rho-gsf}), (\ref{eq:gi-delta-sigma-gsf}) and~(\ref{eq:gi-delta-p-gsf}), respectively, are related as follows: \begin{equation} \widehat{\delta p} =\left[c_{_{\rm S}}^{2}\; \widehat{\delta \rho} +\left(3\, H\right)\; \left(c_{_{\rm S}}^{2}-c_{_{\rm A}}^{2}\right)\; \widehat{\delta \sigma}\right]. \end{equation} This relation, in turn, allows us to write $\delta p^{_{\rm NA}}$ as \begin{equation} \delta p^{_{\rm NA}} =\left(\widehat{\delta p}- c_{_{\rm A}}^2\, \widehat{\delta\rho}\right) = \left(c_{_{\rm S}}^{2} - c_{_{\rm A}}^{2}\right)\, \left[\widehat{\delta \rho} + \left(3\, H\right)\; \widehat{\delta \sigma}\right]. \end{equation} Also, the first order Einstein's equations~(\ref{eq:fo-ee-00}) and~({\ref{eq:fo-ee-0i}) can be combined to arrive at \begin{equation} \left[\widehat{\delta \rho} + \left(3\, H\right)\; \widehat{\delta \sigma}\right] =\left(\frac{1}{4\, \pi\, G\, a^{2}}\right)\; \nabla^{2}\Phi.\label{eq:c12} \end{equation} Evidently, these last two equations immediately lead to the expression~(\ref{eq:delta-p-na-gsf}) for $\delta p^{_{\rm NA}}$. The reason for the appearance of the quantity $c_{_{\rm S}}^2$ in this expression for $\delta p^{_{\rm NA}}$ can be understood as follows. Let us work in a gauge wherein $\delta \phi=0$, a choice of coordinates that is often referred to as the comoving gauge~\cite{reviews}. In such a gauge, no flux of energy arises, i.e. $^{_{\rm (C)}}\delta \sigma=0$ [cf. Eq.~(\ref{eq:delta-sigma-gsf})], and we have used the super-script {\scriptsize ${\rm (C)}$} to denote the fact that we are working in the comoving gauge. It is clear from the expressions~(\ref{eq:delta-rho-gsf}) and~(\ref{eq:delta-p-gsf}) for $\delta \rho$ and $\delta p$ that, in the comoving gauge, we have \begin{equation} ^{_{\rm (C)}}\delta p = c_{_{\rm S}}^{2}\; ^{_{\rm (C)}}\delta\rho. \end{equation} Therefore, the non-adiabatic pressure perturbation $\delta p^{_{\rm NA}}$ in this gauge is given by \begin{equation} \delta p^{_{\rm NA}} = \left[{}^{_{\rm (C)}}\delta p - c_{_{\rm A}}^{2}\; ^{_{\rm (C)}}\delta\rho\right] =\left(c_{_{\rm S}}^{2} - c_{_{\rm A}}^{2}\right)\;\, ^{_{\rm (C)}}\delta\rho \end{equation} and, since $^{_{\rm (C)}}\delta \sigma=0$, the relation~(\ref{eq:c12}) leads to (\ref{eq:delta-p-na-gsf}), as required. \section{Condition for vanishing non-adiabatic pressure perturbation}\label{sec:condition} In Fourier space, $\nabla^{2}\Phi\propto \left(k^{2}\;\Phi_{k}\right)$. It is then clear from Eq.~(\ref{eq:delta-p-na-gsf}) that the non-adiabatic pressure perturbation $\delta p^{_{\rm NA}}$ will vanish in the super Hubble limit (i.e. as $k\to 0$) for all scalar fields~\cite{bep-shs,christopherson-2009}. This occurs, for instance, when the modes are well outside the Hubble radius during the inflationary epoch~\cite{bep-shs}. It is interesting to enquire whether the non-adiabatic pressure perturbation $\delta p^{_{\rm NA}}$ vanishes {\it identically}\/ for any scalar field model. In such a case, the scalar field will behave {\it exactly}\/ like a perfect fluid described by an equation of state, at least at the first order in the perturbations. It is evident from Eq.~(\ref{eq:delta-p-na-gsf}) that such a behavior will be possible if and only if $c_{_{\rm S}}^2 = c_{_{\rm A}}^{2}$. Based on this condition, we shall now arrive at the corresponding condition on the Lagrangian density describing the scalar field. Using Eq.~(\ref{eq:ce}) that describes the conservation of energy, and the expressions~(\ref{eq:rho-p-gsf}) for the background energy density and pressure associated with the scalar field, the adiabatic speed of the perturbations~$c_{_{\rm A}}^{2}$ can be expressed as \begin{equation} c_{_{\rm A}}^{2} =\left(\frac{\dot p}{\dot \rho}\right) =-\left(\frac{\dot p}{\left(3\, H\right)\, \left(\rho+p\right)}\right) = -\left(\frac{{\ddot \phi}\,\left(\partial {\cal L}/\partial X\right) + \left(\partial {\cal L}/\partial \phi\right)}{\left(3\, H\, {\dot \phi}\right)\, \left(\partial {\cal L}/\partial X\right)}\right). \end{equation} The equation of motion~(\ref{eq:em-gsf}) governing the field, then allows us to arrive at the following expression for the adiabatic speed of the perturbations: \begin{eqnarray} c_{_{\rm A}}^{2} &=&-\left(\left(3\, H\, {\dot \phi}\right)\, \left(\frac{\partial {\cal L}}{\partial X}\right)\; \left[\left(\frac{\partial {\cal L}}{\partial X}\right) + (2\, X)\, \left(\frac{\partial^{2} {\cal L}}{\partial X^{2}}\right)\right]\right)^{-1}\nonumber\\ & &\;\;\times\; \biggl[2\, \left(\frac{\partial {\cal L}}{\partial X}\right)\, \left(\frac{\partial {\cal L}}{\partial \phi}\right) -(3\, H\, {\dot \phi})\, \left(\frac{\partial {\cal L}}{\partial X}\right)^{2} + (2\, X)\, \left(\frac{\partial^{2} {\cal L}}{\partial X^{2}}\right)\, \left(\frac{\partial {\cal L}}{\partial \phi}\right) -(2\, X)\, \left(\frac{\partial {\cal L}}{\partial X}\right)\, \left(\frac{\partial^{2} {\cal L}}{\partial X\;\partial \phi}\right)\biggr].\label{eq:ca2} \end{eqnarray} From the definition~(\ref{eq:cs2}) of $c_{_{\rm S}}^{2}$, it is then straightforward to show that the condition $c_{_{\rm A}}^{2} = c_{_{\rm S}}^{2}$ implies that \begin{equation} \left(\frac{\partial {\cal L}}{\partial X}\right)\, \left(\frac{\partial {\cal L}}{\partial \phi}\right) + X\; \left(\frac{\partial^{2} {\cal L}}{\partial X^{2}}\right)\, \left(\frac{\partial {\cal L}}{\partial \phi}\right) -X\; \left(\frac{\partial {\cal L}}{\partial X}\right)\; \left(\frac{\partial^{2} {\cal L}}{\partial X\;\partial \phi}\right)= 0.\label{eq:condition} \end{equation} This condition can be further simplified to be \begin{equation} \frac{\partial}{\partial X}\left[\left(\frac{1}{Y}\right)\; \left(\frac{\partial {\cal L}}{\partial \phi}\right)\right]= 0,\label{eq:condition-sv} \end{equation} where we have defined $Y$ as \begin{equation} Y = X\;\left(\frac{\partial {\cal L}}{\partial X}\right). \end{equation} As we had mentioned before, based on the analogy with fluids, we shall refer to scalar fields that satisfy the condition~(\ref{eq:condition-sv}) as perfect scalar fields. Interestingly, the condition~(\ref{eq:condition}) above has been arrived at earlier while considering a completely different problem involving the stationary configurations of scalar fields~\cite{akhoury-2009}}. This seems to indicate some relation between stationarity and a vanishing non-adiabatic pressure perturbation, which invites further investigation. It is straightforward to show that, if a given Lagrangian density, say, ${\cal L}_{1}(X,\phi)$, satisfies the condition~(\ref{eq:condition-sv}), then the Lagrangian densities, say, ${\cal L}_{2}(X,\phi)$ and ${\cal L}_{3}(X,\phi)$, that are related to ${\cal L}_{1}(X,\phi)$ as follows: \begin{equation} {\cal L}_{2}(X,\phi) = C_{1}\; \exp\, \left[C_{2}\, {\cal L}_{1}(X,\phi)\right] \qquad{\rm and}\qquad {\cal L}_{3}(X,\phi) = C_{3}\; {\rm log}\,\left[C_{4}\, {\cal L}_{1}(X,\phi)\right], \label{eq:rbl} \end{equation} where $C_{1}$, $C_{2}$, $C_{3}$ and $C_{4}$ are constants, also satisfy the same condition. Though these three Lagrangian densities satisfy the condition~(\ref{eq:condition-sv}), it may be worthwhile to note that, they, in fact, lead to different evolution equations for the background as well as the perturbations. \section{Examples of perfect scalar fields}\label{sec:examples} In this section, we shall construct examples of perfect scalar fields that satisfy the condition~(\ref{eq:condition-sv}). We shall first construct purely kinetic models that do not depend on the scalar field directly, and then consider models that also seemingly depend on the scalar field. \subsection{Pure kinetic models} Note that any Lagrangian density~${\cal L}$ that does not explicitly depend on the scalar field~$\phi$, i.e. when $\left(\partial {\cal L}/\partial \phi\right) = 0$, naturally satisfies the condition~(\ref{eq:condition-sv}). These Lagrangian densities are often referred to as pure kinetic models. It is well known that pure kinetic models can be described as barotropic, perfect fluids which, by definition, do not possess any non-adiabatic pressure perturbation (in this context, see Refs.~\cite{tejedor-2005,bertacca-2008}). Such models were originally considered in the context of inflation~\cite{ki}, and they have been resurrected more recently as a potential candidate of dark energy~\cite{de-reviews}. We shall now discuss a couple of specific examples of these models. \subsubsection{A scalar field mimicking a perfect fluid with a constant equation of state} Consider a Lagrangian density of the following form: \begin{equation} {\cal L}(X) = V_{0}\; X^{\alpha},\label{eq:pf} \end{equation} where $V_{0}$ is a constant potential, and $\alpha$ is a real number. Such a Lagrangian density clearly satisfies the condition~(\ref{eq:condition-sv}). The background energy density and pressure corresponding to this Lagrangian density can be obtained to be [cf. Eq.~(\ref{eq:rho-p-gsf})] \begin{equation} \rho = \left(2\, \alpha - 1\right)\, V_{0}\; X^{\alpha}\qquad{\rm and}\qquad p = V_{0}\; X^{\alpha}, \end{equation} so that the resulting equation of state $w \equiv (p/\rho)$ is a constant, and is given by \begin{equation} w = \left(\frac{1}{2\, \alpha - 1}\right). \end{equation} In other words, the scalar field described by the Lagrangian density~(\ref{eq:pf}) essentially behaves like a perfect fluid with a constant equation of state. It is also useful to note that, in such a case, the two speeds of sound simplify to [cf. Eqs.~(\ref{eq:cs2}) and (\ref{eq:ca2})]: $c_{_{\rm A}}^{2} = c_{_{\rm S}}^{2} = w$. \subsubsection{Models that behave as the Chaplygin gas} Consider a tachyon that is governed by the Lagrangian density \begin{equation} {\cal L}(X) = -\left(V_{0}\; \sqrt{1 - 2\, X}\right),\label{eq:tm-cg} \end{equation} where $V_{0}$ is a constant potential. Needless to add, this Lagrangian density also satisfies the condition~(\ref{eq:condition-sv}). The background energy density and pressure corresponding to the above Lagrangian density are given by \begin{equation} \rho = \left(\frac{V_{0}}{\sqrt{1 - 2\, X}}\right)\qquad{\rm and}\qquad p = -\left(V_{0}\, \sqrt{1 - 2X}\right), \end{equation} so that we can write \begin{equation} p = -\left(\frac{V_{0}^{2}}{\rho}\right). \end{equation} This is the equation of state that describes the Chaplygin gas~\cite{kamenshchik-2001,bento-2002,chimento-2004,scherrer-2004} and, in such a case, we obtain that \begin{equation} c_{_{\rm A}}^{2} = c_{_{\rm S}}^{2} = \left(1-2\, X\right)=-w. \end{equation} Upon using the conservation equation~(\ref{eq:ce}) for the background energy density, it can be readily shown that the equation of state parameter of the Chaplygin gas evolves as a function of the scale factor in the following fashion: \begin{equation} w(a) = -\left(\frac{V_{0}^{2}}{V_{0}^{2}+({\cal A}/a^{6})}\right), \end{equation} where ${\cal A}$ is a positive constant. A more generic form of the Lagrangian density~(\ref{eq:tm-cg}) that satisfies the condition~(\ref{eq:condition-sv}) is given by \begin{equation} {\cal L}(X) = -\left(V_{0}\; \left[1 - \left(X/V_{0}\right)^{[(1 + \alpha)/2\,\alpha]}\right]^{[\alpha/(1 + \alpha)]}\right),\label{eq:tm-gcg} \end{equation} where $V_{0}$ is again a constant, while $\alpha$ is a real number. It is straightforward to show that, for this Lagrangian density, the homogeneous energy density and pressure are given by \begin{equation} \rho=\left(V_{0}\; \left[1 - \left(X/V_{0}\right)^{[(1+ \alpha)/2\, \alpha]}\right]^{-[1/(1 + \alpha)]}\right) \qquad{\rm and}\qquad p=-\left(V_{0}\; \left[1 - \left(X/V_{0}\right)^{[(1 + \alpha)/2\,\alpha]}\right]^{[\alpha/(1 + \alpha)]}\right), \end{equation} so that, we have \begin{equation} p = -\left(\frac{V_{0}^{(1 + \alpha)}}{\rho^{\alpha}}\right), \end{equation} which is the equation of state of the generalized Chaplygin gas. Over the last few years, such models have been considered in the literature as possible candidates for dark energy~\cite{bento-2002,chimento-2004,scherrer-2004}. In this case, the equation of state parameter $w$ evolves as \begin{equation} w(a) = -\left(\frac{V_{0}^{(1 + \alpha)}}{V_{0}^{(1 + \alpha)} +\left({\cal A}/a^{[3\, \left(1+\alpha\right)]}\right)}\right), \end{equation} where ${\cal A}$ is again a positive constant. Also, we find that \begin{equation} c_{_{\rm A}}^{2} = c_{_{\rm S}}^{2} = \alpha\; \left[1 - \left(X/V_{0}\right)^{[(1+ \alpha)/2\,\alpha]}\right]. \end{equation} \subsubsection{Another model} Now, consider a Lagrangian density of the following form: \begin{equation} {\cal L}(X) = -V_{0}\left(\sqrt{X}- 1\right)^{-\alpha},\label{eq:dm-de} \end{equation} where, again, $V_{0}$ is a constant, and $\alpha$ is a real number. The background energy density and pressure corresponding to the above Lagrangian density are found to be \begin{equation} \rho = V_{0}\, \left[(1+\alpha)\, \sqrt{X} -1\right]\; \left(\sqrt{X}- 1\right)^{-(1+\alpha)} \qquad{\rm and}\qquad p = -V_{0}\, \left(\sqrt{X}- 1\right)^{-\alpha}.\label{eq:dm-de-rho-p} \end{equation} In such a case, it can be shown that the equation of state parameter $w$ evolves with the scale factor as \begin{equation} w(a)=-\left(\frac{1}{\left(1+\alpha\right)+\left(\alpha/{\cal A}\right)\, a^{-\left[3/(1+\alpha)\right]}}\right),\label{eq:w-another-model} \end{equation} where, as before, ${\cal A}$ is a positive constant, while \begin{equation} c_{_{\rm A}}^{2} = c_{_{\rm S}}^{2} =\left(\frac{1-\sqrt{X}}{\left(1+\alpha\right)\, \sqrt{X}}\right). \end{equation} Note that the expression~(\ref{eq:w-another-model}) for $w(a)$ implies that, for positive $\alpha$, $w(a)$ evolves from zero to an asymptotic value of $-(1 + \alpha)^{-1}$. For an appropriate choice of the constants ${\cal A}$ and $\alpha$, it is then possible to achieve $w(a)$ at the present epoch to be less than $-(1/3)$. Hence, the above Lagrangian density may be considered as a possible model of dark energy. \subsection{Masquerading scalar fields} We find that we can construct three types of models that {\it seemingly}\/ depend on the scalar field and also satisfy the condition~(\ref{eq:condition-sv}). The first type are models wherein the Lagrangian density of the scalar field can be expressed as a sum of the kinetic and the potential terms as follows: \begin{equation} {\cal L}(X,\phi) = f(X)-V(\phi).\label{eq:one} \end{equation} In such a situation, the condition~(\ref{eq:condition-sv}) immediately restricts the function $f(X)$ to be \begin{equation} f(X)=\alpha\; {\rm log}\, \left(X/X_{0}\right),\label{eq:two} \end{equation} where $\alpha$ and $X_{0}$ are positive constants. Moreover, the energy density and pressure associated with the homogeneous scalar field are given by \begin{equation} \rho=\,\alpha\, \left[2-{\rm log}\, \left(X/X_{0}\right)\right]+V(\phi) \qquad{\rm and}\qquad p =\alpha\; {\rm log}\, \left(X/X_{0}\right)-V(\phi), \end{equation} and we find that the equation of state parameter evolves as \begin{equation} w(a) = -1 +\left(\frac{2\, \alpha}{{\cal A} - 6\, \alpha\, \log\, a}\right), \end{equation} where ${\cal A}$ is a suitably chosen positive constant. Also, we obtain that $c_{_{\rm A}}^{2} = c_{_{\rm S}}^{2} =-1$. The second type of models are wherein we can write the Lagrangian density as a product of the kinetic and the potential terms in the following fashion~\cite{ki,chimento-2004,scherrer-2004}: \begin{equation} {\cal L}(X,\phi) = f(X)\, V(\phi).\label{eq:ttm} \end{equation} In such cases, the condition~(\ref{eq:condition-sv}) restricts the function~$f(X)$ to be $X^{\alpha}$, where $\alpha$ is a real number. The smooth energy density and pressure associated with this Lagrangian density can be obtained to be \begin{equation} \rho=(2\, \alpha-1)\, \left[X^{\alpha}\, V(\phi)\, \right] \qquad{\rm and}\qquad p=\left[X^{\alpha}\,V(\phi)\, \right]. \end{equation} The resulting equation of state is evidently a constant, and is given by \begin{equation} w = \left(\frac{1}{2\, \alpha - 1}\right). \end{equation} Obviously, this example is a more general case of the Lagrangian density~(\ref{eq:pf}) we had discussed before. As in the earlier example, the adiabatic and the effective speeds of sound associated with the perturbations reduce to \begin{equation} c_{_{\rm A}}^{2} = c_{_{\rm S}}^{2} = \left(\frac{1}{2\,\alpha-1}\right)=w. \end{equation} Note that the two Lagrangian densities that we have considered in this sub-section until now can be related to each other through the relations~(\ref{eq:rbl}). The third type of scalar fields that we shall consider are those that are described by the Lagrangian density \begin{equation} {\cal L}(X,\phi) = {\cal B}\, \exp\, \left[X^{\alpha}\, V(\phi)\right], \label{eq:gm} \end{equation} where ${\cal B}$ is a negative constant, and $\alpha$ is a real number. Clearly, as the previous example, this Lagrangian density too will satisfy the condition~(\ref{eq:condition-sv}), courtesy of the relations~(\ref{eq:rbl}). We obtain the homogeneous energy density and pressure associated with the above Lagrangian density to be \begin{equation} \rho= {\cal B}\,\, \left[2\,\alpha\, X^{\alpha}\, V(\phi)-1\right]\; \exp\, \left[X^{\alpha}\, V(\phi)\right]\, \qquad{\rm and}\qquad p={\cal B}\, \exp\, \left[X^{\alpha}\,V(\phi)\,\right]. \end{equation} Also, we find that \begin{equation} c_{_{\rm A}}^{2} = c_{_{\rm S}}^{2} = \left(\frac{1}{2\,\alpha\, \left(1+V(\phi)\, X^{\alpha}\right)-1}\right), \end{equation} so that we can write \begin{equation} c_{_{\rm S}}^{2} = \left(\frac{w}{2\,\alpha\, w + 1}\right). \end{equation} Note that, if $\alpha > (3/2)$, then $c_{_{\rm S}}^{2}$ is a positive definite quantity, and $w$ can be less than $-(1/3)$. Though the above three examples behave as though they explicitly depend on the scalar field, we find that they can be turned into pure kinetic models by a suitable redefinition of the field. For instance, the transformation \begin{equation} \chi(\phi)=\int\; d\phi\; \left[\frac{V(\phi)}{V_{0}}\right]^{(1/2\,\alpha)}, \end{equation} reduces the Lagrangian density~(\ref{eq:ttm}) to the example~(\ref{eq:pf}). Evidently, such a transformation will also convert the Lagrangian density~(\ref{eq:gm}) into a purely kinetic form. Moreover, we find that the following transformation \begin{equation} \chi(\phi)=\int\; d\phi\; \exp-\left[V(\phi)/2\,\alpha\right], \end{equation} removes the explicit dependence on the scalar field in the first example above [cf.~Eqs.~(\ref{eq:one}) and (\ref{eq:two})]. In other words, classically, the three Lagrangian densities that we have presented in this sub-section are essentially pure kinetic models that masquerade as though they depend on the scalar field. The fact that field redefinitions can relate different, but dynamically equivalent, Lagrangian densities for scalar fields has been noticed earlier (in this context, see, in particular, Ref.~\cite{bertacca-2008}). However, we should add that it is not clear whether such a behavior will be preserved when the quantum nature of the scalar fields are taken into account, as it is required, say, in the context of inflation. In this section, we had presented a few specific examples of scalar fields that satisfy the condition~(\ref{eq:condition-sv}). It should be noted that the most general solution to this equation can be expressed in terms of the product $\left[X\;f(\phi)\right]$, where $f(\phi)$ is an arbitrary function of the scalar field (for details, see Ref.~\cite{akhoury-2009}). \section{An interesting implication for inflation}\label{sec:ii} We find that the perfect scalar fields that satisfy the condition~(\ref{eq:condition-sv}) behave in a particular fashion when they are treated as the inflaton. We shall now turn to discuss this property. It is well known that, in the slow roll inflationary scenario involving a single scalar field, the amplitude of the curvature perturbation approaches a constant value soon after the modes leave the Hubble radius (see, any of the standard texts~\cite{texts} or reviews~\cite{reviews}). However, this is not true when there are one or more brief periods of deviation from slow roll inflation\footnote{Actually, it could even be a short period of departure from inflation. But, the deviation from slow roll {\it has}\/ to be brief, since a prolonged deviation may not allow inflation to restart.}. In such cases, the amplitude of the curvature perturbation over a suitable range of modes (which leave the Hubble radius just before the deviation from slow roll) can be enhanced or suppressed at super Hubble scales when compared to their value at Hubble exit. It can be shown that this behavior essentially arises due to the fact that, in such scenarios, the intrinsic entropy perturbation (i.e. the non-adiabatic pressure perturbation) grows rapidly around the period of fast roll at super Hubble scales~\cite{bep-shs}. Since the non-adiabatic pressure perturbation vanishes identically, the phenomenon described above can not occur in models which satisfy the condition~(\ref{eq:condition-sv}). It is then possible that the demand that the non-adiabatic pressure perturbation vanishes exactly is so restrictive that the perfect scalar fields will not allow a brief period of departure from slow roll inflation. This expectation indeed turns out to be true and, in what follows, we shall outline its proof. Recall that the first slow roll parameter~$\epsilon$ is defined as~\cite{texts,reviews} \begin{equation} \epsilon = -\left(\frac{\dot H}{H^2}\right)=\left(\frac{3\, (1 + w)}{2}\right), \end{equation} and inflation corresponds to the epoch wherein $\epsilon<1$ or, equivalently, $w< -(1/3)$. The time derivative of the quantity~$\epsilon$ is given by \begin{equation} {\dot \epsilon} = \left(\frac{9\, H}{2}\right)\, \left(1+w\right)\, \left(w-c_{_{\rm A}}^2\right) =\left(\frac{9\, H}{2}\right)\, \left(1+w\right)\, \left(w-c_{_{\rm S}}^2\right), \label{eq:epsilon-dot} \end{equation} where, in arriving at the last equality, we have set $c_{_{\rm A}}^2 = c_{_{\rm S}}^{2}$, as it is in the case of perfect scalar fields. The scalar power spectrum is determined by the amplitude of the curvature perturbations when the modes are well outside the Hubble radius during the inflationary epoch~\cite{texts,reviews}. Also, these curvature perturbations are evolved from the sub Hubble to the super Hubble scales by imposing standard initial conditions (viz. the Bunch-Davies conditions) when the modes are well inside the Hubble radius. It is the quantity $c_{_{\rm S}}$ that describes the speed of propagation of the curvature perturbations. Hence, $c_{_{\rm S}}$ has to be a real quantity, if it has to be ensured that the curvature perturbations do not grow exponentially at sub Hubble scales, so that the Bunch-Davies initial conditions can be imposed on them~ (see Ref.~\cite{ki}; in this context, also see Ref.~\cite{shanki-2009}). So, during inflation, in addition to $w<-(1/3)$, we require that $c_{_{\rm S}}^{2}>0$. It is then evident from the last equality in the expression~(\ref{eq:epsilon-dot}) above that \begin{equation} {\dot \epsilon}<0. \end{equation} In other words, $\epsilon$ will {\it always}\/ be a monotonically {\it decreasing}\/ function of time when perfect scalar fields drive inflation. Therefore, clearly, once slow roll inflation has commenced, these models will not allow a period of deviation from slow roll inflation. In fact, terminating inflation becomes a problem in such cases, and one will have to invoke an additional scalar field to exit inflation. We should point out that, amongst the various examples of perfect scalar fields that we have discussed, only the models described in Eqs.\~(\ref{eq:tm-cg}), (\ref{eq:tm-gcg}) and~(\ref{eq:gm}) satisfy both the conditions $w<-(1/3)$ and $c_{_{\rm S}}^{2}>0$ to act as a inflaton. It has been recognized that certain features in the primordial spectrum lead to a better fit to the cosmic microwave background data than the conventional, featureless, power law, primordial scalar perturbation spectrum (in this context, see, for example, Refs.~\cite{fips}, and the long list of references therein). Interestingly, a short epoch of deviation from slow roll inflation is essential for generating features in the inflationary, scalar perturbation spectrum. In particular, it seems mandatory that the first slow roll parameter $\epsilon$ has to rise and fall if features are to be produced in the scalar power spectrum\footnote{For example, it is known that the spectrum can remain featureless even if the second slow roll parameter is large for a short duration of time (see, for instance, Refs.~\cite{kinney}).}. However, as we discussed above, it is not possible to achieve such a behavior for $\epsilon$ in perfect scalar field models. This, in turn, implies that scalar fields with a vanishing non-adiabatic pressure perturbation, if they are treated as the inflaton, can not lead to features in the scalar perturbation spectrum! \section{Concluding remarks}\label{sec:cr} In this paper, we have obtained a condition on the Lagrangian density of an arbitrary scalar field under which the intrinsic, non-adiabatic pressure perturbation associated with the scalar field vanishes identically, at the first order in the perturbations. Motivated by the analogy with fluids, we have termed these fields as perfect scalar fields. Scalar field models that depend only on the kinetic energy are known to contain no non-adiabatic pressure perturbation. In addition to discussing a couple of such examples, we have presented a few models which seemingly depend on the scalar field explicitly, and possess no non-adiabatic pressure perturbation. But, we find that all such models that we were able to construct could be reduced to purely kinetic models by a redefinition of the field. In fact, it can be proved that all Lagrangian densities that satisfy the condition~(\ref{eq:condition-sv}) can be reduced to a purely kinetic form (in this context, see the discussion following Eq.~(2.22) in Ref.~\cite{akhoury-2009}). Interestingly, we have also shown that, if perfect scalar fields are used to drive inflation, then they can not lead to features in the scalar power spectrum. Note that, throughout the paper, we had restricted our attention to the first order in the perturbations. We should point out that, since all the scalar fields that satisfy the condition~(\ref{eq:condition-sv}) can be reduced to pure kinetic models, such perfect scalar fields will not contain any non-adiabatic pressure perturbation at the higher orders in the perturbations either (in this context, see Refs.~\cite{sopt}). In other words, these scalar fields are truly perfect. \section*{Acknowledgments} LS and SU wish to thank the Inter University Centre for Astronomy and Astrophysics, Pune, India, and the Harish-Chandra Research Institute, Allahabad, India, for hospitality, where this work was initiated and completed, respectively. LS also wishes to thank Will Kinney and Jerome Martin for discussions, and Hermano Velten for drawing attention to a few relevant references.
\section{Introduction} \label{intro} Diophantine equations of the form $px^2+c=y^p$, where $c$ is a nonzero integer and $p$ is an odd prime, have been studied by several authors. When $c=2^n$, the case $p=3$ was solved by Rabinowitz in \cite{RAB}, while Le dealt with the case $p>3$ in \cite{LE}. The case $c=3^n$ was considered by Abu Muriefah in \cite{AMT}. Cao \cite{CAO} treated the cases $c=1$ and $c=4^n$ (see also \cite{AMI}, \cite{ADA}, \cite{BROWN}, \cite{COHN} for closely related results). We should moreover mention that the equation has no solution in positive integers $x,y$ when $c=-1$, as can be inferred from the work of Nagell \cite{NAG} and Cao \cite{CAO1}. The case when $c=q^{2n}$, where $q$ is an odd prime, was studied in the recent paper \cite{AM} of Abu Muriefah. Let us first note that, for fixed $n$ and $p\geq 5$, the Diophantine equation $px^2+q^{2n}=y^p$ and, more generally, the Diophantine equation $pX^2+Y^{2n}=Z^p$, have at most finitely many solutions $(x,q,y)$ and $(X,Y,Z)$, respectively. Indeed, in this case, $\frac{1}{2}+\frac{1}{2n}+\frac{1}{p}<1$ and the claim follows from Theorem 2 of \cite{DaGra}. A main result in the aforementioned paper \cite{AM}, namely, Theorem 3.1, states that the equation $5x^2+q^{2n}=y^5$, has two families of solutions given by $y=\phi_{3k},\phi_{3k+1}$ (or $\psi_{3k+1},\psi_{3k+2})$, \ $k>1$, where $\phi_{k}$ (respectively $\psi_{k}$) is the $k$th term of the Fibonacci sequence (respectively the Lucas sequence). However, straightforward computations show that the only Fibonacci or Lucas number $y<1000$ satisfying the title equation when $p=5$ is $y=21$ with $x=410$, $q=1801$, $n=1$, and, further, in $5\times183630^{2}+160201^{2}=181^{5}$, $160201$ is prime and $181$ is neither a Fibonacci nor a Lucas number. The same Theorem 3.1 of \cite{AM} states that, if $p>3$ is a prime $\not\equiv 7\pmod 8$ and $q$ is another odd prime, then there are no integer solutions $(x,y,n)$ to the equation $px^2+q^{2n}=y^p$ with $(x,y)=1$. The proof of Theorem 3.1 in \cite{AM}, just before its end, contains an obvious, non-rectifiable error, at case 2, when $-16apb^2$ is set equal to $-16apq^{2m}$ although $b=\pm q^j$ with $0\leq j<m$. That the said proof is erroneous is also pointed out by P.G.~Walsh in his review \cite{Walsh} of \cite{AM}. One of our aims in the present paper is to prove Abu Muriefah's assertion when $p\equiv 3\pmod 8$; see Theorem \ref{th p equiv 3 mod 8}. Our proof, rather than rectifying Abu Muriefah's argument (this is probably impossible), goes through totally different lines. Unfortunately, our arguments cannot be extended to the case $p\equiv 1\pmod 4$. As we revisited the title equation, we further discovered some new results, like Theorems \ref{th on main eq}, \ref{th gen exp eq} and \ref{th with y=q} which, we believe, merit one's attention. Moreover, since the powerful techniques of sections \ref{section exp eq with y prime} and \ref{section general exp eq} are also applicable (after the appropriate modifications) to Diophantine equations other than the ones treated in this paper, we thought it useful to expose them in some detail, enough for the reader to profit from them. As we stated above, one of the main results of this paper is the following \begin{theorem} \label{th on main eq} Let $q$ be an odd prime. If either $q\not\equiv 1\pmod{600}$ or $q \leq 3\cdot 10^9$, then there is no integer solution $(x,y,n)$ to the equation \begin{equation} \label{eq initial} 5x^2+q^{2n} = y^5\,,\quad x,y,n >0\,. \end{equation} Otherwise, there exists at most one integer solution $(x,y,n)$ and if it actually exists, then it must satisfy the following conditions: \\ (i) $n<820$ and ${\mathrm{gcd}}(n,2\cdot 3\cdot 5\cdot 7\cdot 11\cdot 13)=1$. \\ (ii) There exists an integer $v$ such that $x=10v(80v^{4}-40v^{2}+1)$, $ y=20v^{2}+1$, $ q^{n}=2000v^{4}-200v^{2}+1$. \end{theorem} {\em Remark}. If the prime $q$ is of the form $q=2000v^4-200v^2+1$ (the first few primes of this shape are 1801, 160201, 1245001, 4792201, 8179201), then $(x,y,n)=(10v(80v^4-40v^2+1),20v^2+1,1)$ is a solution to (\ref{eq initial}) and, according to the theorem, this is the only one (with $x,y>0$). We have not been able, however, to find a prime $q$ such that the corresponding equation (\ref{eq initial}) has a solution with $n>1$. Therefore, we state the following \begin{conjecture} If the prime $q$ is not of the form $q=2000v^4-200v^2+1$, then the equation (\ref{eq initial}) has no solutions. \end{conjecture} The proof of Theorem \ref{th on main eq} follows from a straightforward combination of Corollary \ref{proposition on main eq} and Theorem \ref{th with y=q}, a second main result of our paper. In turn, the proof of Theorem \ref{th with y=q} relies on a third main result, namely, Theorem \ref{th gen exp eq} concerning the equation $5x^2-4=y^n$ which is interesting for its own sake. Indeed, in recent years, important papers are devoted to equations of the form $x^2+C=y^n$. One main strategy for attacking such equations is based on the so called {\em modular method} which has been successfully applied in quite a number of cases; see Chapter 15 (by S.~Siksek) in H.~Cohen's book \cite{Cohen}, the survey article \cite{FSAbu} and \cite{BMSII} and the references therein. For our equation $5x^2-4=y^n$, the existence of the trivial solution $(x,y)=(1,1)$ makes the modular approach unsuccessful and prevents us from giving the complete solution $(x,y,n)$. Thus, our Theorem \ref{th gen exp eq} offers only a partial result which, at present, seems to be best possible. \section{The Diophantine equation $px^2+q^{2n}=y^p$} The main results of this section are Proposition \ref{proposition general Muriefah} and Theorem \ref{th p equiv 3 mod 8} \begin{proposition} \label{proposition general Muriefah} Let $p>3$ be a prime number $\not\equiv 7 \pmod 8$ and let $q$ be an odd prime number. If $x,y,n$ are positive integers with $x,y$ relatively prime such that \begin{equation} \label{eq 3.1} px^2+q^{2n}=y^p\,, \quad (x,y)=1, \ n>0\,, \end{equation} then there exists a rational integer $a$ such that \begin{equation} \label{q^n} \pm q^n = \sum_{i=0}^{(p-1)/2} \bincf{p}{2i+1}a^{p-2i-1}(-p)^{(p-2i-1)/2} \end{equation} and \begin{equation} \label{x} x = \sum_{i=0}^{(p-1)/2} \bincf{p}{2i}a^{p-2i}(-p)^{(p-2i-1)/2} \end{equation} \end{proposition} {\bf Proof}.\: The condition $(x,y)=1$ implies $p\neq q$ and the condition $p\not\equiv 7\pmod 8$ implies that $y$ is odd. We work in the imaginary quadratic field $K={\mathbb Q}({\omega})$, where ${\omega}=\sqrt{-p}$. Equation (\ref{eq 3.1}) factorizes as $({\omega} x+q^n)(-{\omega} x+q^n)=y^p$ and, trivially, the factors in the left-hand side are relatively prime. This implies an ideal equation $\ideal{{\omega} x+q^n}={\bf I}^p$, where ${\bf I}$ is an integral ideal of $K$. Since the ideal-class number of $K$ is strictly less than $p$ (see page 199 of \cite{Fai}), the above ideal equation implies that ${\bf I}$ is a principal ideal, therefore we obtain the equation \begin{equation} \label{xomega+q^n} x{\omega} +q^n = {\alpha} ^p\,,\quad {\alpha}=a{\theta}+b\,, \quad {\theta} =\begin{cases} {\omega} & \mbox{if $p\equiv 1,5\pmod 8$} \\ \frac{1+{\omega}}{2} & \mbox{if $p\equiv 3\pmod 8$} \end{cases} \end{equation} for some rational integers $a$ and $b$. In case $p\equiv 3\pmod 8$ we write the above equation as $q^n-x +2x{\theta}=(b+a{\theta})^p$, implying $(b+a{\theta})^p\equiv\mbox{$0$ or $1\pmod 2$}$. From this, we see that $a$ cannot be odd. For, otherwise, we would have ${\theta}^p$ or $(1+{\theta})^p\equiv\mbox{$0$ or $1\pmod 2$}$. But we easily check that, for $k\not\equiv 0\pmod 3$, it is true that ${\theta}^k, (1+{\theta})^k\equiv \mbox{${\theta}$ or $1+{\theta}\pmod 2$}$, a contradiction. Therefore, $a$ in (\ref{xomega+q^n}) is even and (\ref{xomega+q^n}) is equivalent to the simpler equation \begin{equation} \label{xomega+q^n simpler} x{\omega} +q^n = {\alpha} ^p\,,\quad {\alpha}=a{\omega}+b\,, \end{equation} for some rational integers $a$ and $b$ of opposite parity since $y=a^2p+b^2$ is odd. Also, it is easy to see that $(pa,b)=1$. If we put \[ {\beta}=-\bar \alpha\,, \] then \[ x = \frac{{\alpha}^p + {\beta}^p}{2{\omega}}\,, \quad \frac{q^n}{b}=\frac{{\alpha}^p-{\beta}^p}{{\alpha}-{\beta}} \] and the fact that $({\alpha}^p-{\beta}^p)/({\alpha}-{\beta})$ is an algebraic integer implies that $b$ divides $q^n$ (in ${\mathbb Z}$), hence $b=\pm q^j$ for some $j\in\{0,1,\ldots,n\}$. At this point we note that the pair $({\alpha},-{\beta})$ is a Lehmer pair for which \[ ({\alpha}^2-{\beta}^2)^2=-16pa^2b^2\,. \] Concerning $j$ appearing in the relation $b=\pm q^j$ (see a few lines above), we distinguish two cases. \\ (\rm{i}) $j>0$. Then, in the terminology of \cite{BHV}, $({\alpha},-{\beta})$ is a $p$-defective pair. By Theorems 1.4 and C of \cite{BHV} it easily follows that $p=5$ is the only possibility. Then, by Theorem 1.3 of \cite{BHV}, either $20a^2=\phi_{k-2{\epsilon}}$ for some $k\geq 3$, or $20a^2=\psi_{k-2{\epsilon}}$ for some $k\neq 1$, where ${\epsilon}\in\{-1,1\}$, $(\phi_n)_{n\geq 0}$ denotes the Fibonacci sequence and $(\psi_n)_{n\geq 0}$ is defined by $\psi_0=2$, $\psi_1=1$ and $\psi_n=\psi_{n-1}+\psi_{n-2}$ for $n\geq 2$. It is easily checked that, for every $n\geq 0$, $\psi_n\not\equiv 0\pmod 5$; therefore the second alternative must be excluded. On the other hand, by Th\'{e}or\`{e}me 1.3 of \cite{BMSIII}, a relation of the form $\phi_k=5z^m$ with $m>1$ and $z>1$ is impossible, which excludes the first alternative as well. \\ (\rm{ii}) $j=0$, so that $b=\pm 1$. Then, equating rational and irrational parts in (\ref{xomega+q^n simpler}), we respectively obtain the relations (\ref{q^n}) and (\ref{x}). \hspace*{1mm} \hfill{$\Box$} \begin{corollary} \label{proposition on main eq} If the integers $x,y,q,n$, where $(x,y)=1$, $q$ is an odd prime and $n\geq1$, satisfy the equation \begin{equation} \label{Abu eq} 5x^{2}+q^{2n}=y^{5} \end{equation} then $(n,6)=1$, $q\equiv 1\pmod{600}$ and there exists an integer $v$ such that \begin{equation}\label{shape of sols} x=10v(80v^{4}-40v^{2}+1)\,,\quad y=20v^{2}+1\,,\quad q^{n}=2000v^{4}-200v^{2}+1\,. \end{equation} \end{corollary} {\bf Proof}.\: Applying Proposition \ref{proposition general Muriefah} with $p=5$ we obtain $\pm q^n=125a^4-50a^2+1$ and $x=5a(5a^4-10a^2+1)$. Obviously, the minus sign in the first equation is rejected and $a$ is $a$ is even. Putting $a=2v$ in these relations we obtain the first and third relation in (\ref{shape of sols}), and then the second relation results immediately. \\ We claim that $n$ is odd. Indeed, otherwise $(a,q^{n/2})$ would be an integral point on the elliptic curve defined by $Y^2=125X^4-50X^2+1$. But this elliptic curve has zero rank and its only rational point is $(X,Y)=(0,\pm 1)$, which forces $q=1$, a contradiction. \\ We also claim that $n$ is prime to 3. Indeed, let us write the third equation (\ref{shape of sols}) as $q^n+4=5w^2$, where $w=20v^2-1$. If $n$ were divisible by 3, then the last equation could be written as $(5^2w)^2=(5q^{n/3})^3+500$, again forcing $q=1$, because it is well known since long (see, for example, Table 8 in \cite{LoFi}) that the only integral solutions $(X,Y)$ to $Y^2=X^3+500$ are $(X,Y)=(5,\pm 25)$. \\ Finally, we show that $q\equiv 1\pmod{600}$. First, we write the third equation (\ref{shape of sols}) as $q^n=200v^2(v^2-1)+1$, which shows that $q^n\equiv 1\pmod{600}$. Let $r$ be the order of $q$ modulo $600$. Then $r$ divides $n$, and since $\varphi(600)=160$ and $n$ is odd, we obtain $r=1$ or $r=5$. The latter case cannot hold, for, otherwise, $5x^{2}=y^{5}-(q^{2n/5})^{5}$, which has no proper solutions by Th\'{e}or\`{e}me 2(2) of \cite{Kraus}. We therefore conclude that $r=1$ and $q\equiv 1\pmod{600}$, as claimed. \hspace*{1mm} \hfill{$\Box$} \begin{theorem} \label{th p equiv 3 mod 8} Let $p,q$ be odd primes with $p\equiv 3\pmod 8$ and $p>3$. Then, the Diophantine equation \begin{equation} \label{eq p equiv 3 mod 8} px^2 +q^{2n} = y^p \end{equation} has no positive integer solutions $(x,y,n)$ with $(x,y)=1$. \end{theorem} {\bf Proof}.\: By the relation (\ref{q^n}) we see that the equation (\ref{eq p equiv 3 mod 8}) implies \[ \pm q^n = \frac{(a{\omega} +1)^p -(a{\omega}-1)^p}{2}\,. \] Let us consider the polynomial \[ f(x):=\frac{({\omega} x +1)^p -({\omega} x-1)^p}{2}\,. \] Clearly, this is a polynomial in ${\mathbb Z}[x]$ of degree $p-1$, with leading coefficient $-p^{(p+1)/2}$ and constant term $1$. \begin{quote} {\em First Claim}: The polynomial $f(x)$ factorizes over ${\mathbb Q}[x]$ into two relatively prime polynomials $f_1(x),f_2(x)\in{\mathbb Z}[x]$, each of degree $(p-1)/2$. \end{quote} Proof: Let ${\zeta}$ be a primitive $p$-th root of unity, i.e. a root of the $p$-th cyclotomic polynomial $\Phi(x)=x^{p-1}+\cdots+x+1 = (x^p-1)/(x-1)$. Let also $g$ be a primitive root $\bmod\,p$. Observe that the field $L={\mathbb Q}({\zeta})$ contains ${\omega}$. Indeed, it is a well-known fact that the Gauss sum $G=\sum_{t=1}^{p-1}\jac{t}{p}{\zeta}^t$ satisfies $G^2=(-1)^{(p-1)/2}p$. Therefore, we can assume ${\omega}=G$. Below we will use the well-known fact that, in the field $L$ we have the factorization into prime ideals $\ideal{p}=\ideal{{\lambda}}^{p-1}$, where ${\lambda}=1-{\zeta}$. For ${\beta}\in L$ we will write $\mathrm{w}({\beta})$ to denote $\mathrm{v}_{{\lambda}}({\beta})$, the exponent of ${\lambda}$ in the prime factorization of ${\beta}$; and for $b\in{\mathbb Q}$ we will write $\mathrm{v}(b)$ to denote $\mathrm{v}_{p}(b)$, the exponent of $p$ in the prime factorization of $b$. \\ We have \[ f(x)=\prod_{k=1}^{p-1}\{({\omega} x+1)-{\zeta}^k({\omega} x-1)\}\,, \] from which it follows that the roots of $f(x)$ are exactly the following: \[ \frac{{\omega}}{p}\cdot\frac{1+{\zeta}^k}{1-{\zeta}^k}\,,\quad k=1,\ldots,p-1\,. \] Therefore, \[ f(x)=-p^{(p+1)/2}\prod_{k=1}^{p-1}\left(x- \frac{{\omega}}{p}\cdot\frac{1+{\zeta}^k}{1-{\zeta}^k}\right)\,. \] Let us put now \begin{eqnarray*} f_1(x) & = & p^{(p+1)/4}\prod_{j=0}^{(p-3)/2}\left(x-\frac{{\omega}}{p}\cdot\frac{1+{\zeta}^{g^{2j}}}{1-{\zeta}^{g^{2j}}}\right) \\ f_2(x) & = & -p^{(p+1)/4}\prod_{j=0}^{(p-3)/2}\left(x-\frac{{\omega}}{p}\cdot\frac{1+{\zeta}^{g^{2j+1}}}{1-{\zeta}^{g^{2j+1}}}\right) \end{eqnarray*} so that $f(x)=f_1(x)f_2(x)$. We now show that the polynomials $f_i(x)$ have rational coefficients. \\ The Galois group of the extension ${\mathbb Q}({\zeta})/{\mathbb Q}$ is cyclic generated by the automorphism ${\sigma}$, defined by ${\sigma}({\zeta})={\zeta}^g$. Since ${\omega}\in{\mathbb Q}({\zeta})\setminus{\mathbb Q}$, we must have ${\sigma}({\omega})\neq{\omega}$, therefore, ${\sigma}({\omega})=-{\omega}$. Consequently, for the typical root of $f_1(x)$ we have \begin{equation} \label{action of sigma} {\sigma}\left(\frac{{\omega}}{p}\cdot\frac{1+{\zeta}^{g^{2j}}}{1-{\zeta}^{g^{2j}}}\right) = -\frac{{\omega}}{p}\cdot\frac{1+{\zeta}^{g^{2j+1}}}{1-{\zeta}^{g^{2j+1}}} = \frac{{\omega}}{p}\cdot\frac{1+{\zeta}^{g^{2j'}}}{1-{\zeta}^{g^{2j'}}}\,, \end{equation} where \[ j'\equiv \frac{p+1}{4}+j\pmod{\frac{p-1}{2}} \] and we choose \[ j'=\begin{cases} \frac{p+1}{4}+j & \mbox{if $0\leq j < \frac{p-3}{4}$} \\ \frac{p+1}{4}+j-\frac{p-1}{2} & \mbox{if $\frac{p-3}{4}\leq j\leq\frac{p-3}{2}$} \end{cases}\,, \] so that $j'$ runs (exactly once) through all values $0,1,\ldots,\frac{p-3}{2}$ as $j$ runs through these values. Consequently, the coefficients of the polynomial $f_1(x)$ are fixed by ${\sigma}$, which implies that they belong to ${\mathbb Q}$; and similarly for $f_2(x)$. Actually, the coefficients of $f_1(x),f_2(x)$ are integers and we prove this as follows. \\ First, we show that the absolute value of the constant coefficient of both $f_1(x)$ and $f_2(x)$ is 1. Indeed, let $b_i$ be the constant coefficient of $f_i(x)$. We already know that $b_i\in{\mathbb Q}$. Moreover, multiplying the right equalities (\ref{action of sigma}) for $j=0,\ldots,(p-3)/2$ and then the resulting products in the two sides by $-p^{(p+1)/4}$, we obtain $b_2=b_1$. But, $b_1b_2$ is equal to the constant term of $f(x)$, which is $1$. Therefore $1=b_1b_2=b_1^2$, from which $b_1=b_2=\pm 1$. \\ Let us put now \[ g(x)=x^{p-1}f(\frac{1}{x})\,,\quad g_1(x)=x^{(p-1)/2}f_1(\frac{1}{x})\,,\quad g_2(x)=x^{(p-1)/2}f_2(\frac{1}{x})\,, \] i.e.~these are the reciprocal polynomials of $f(x)$, $f_1(x)$ and $f_2(x)$, respectively. Since $f(x)=f_1(x)f_2(x)$, we also have $g(x)=g_1(x)g_2(x)$. Since the constant term of $f(x)$ is $1$, $g(x)$ has leading coefficient $1$; and since $f(x)$ has integer coefficients, so does $g(x)$. Therefore, the roots of $g(x)$ are algebraic integers. Analogously, the polynomials $g_i(x)$ have leading coefficients equal to $\pm 1$, their coefficients are rational numbers and their roots, being roots of $g(x)$, are algebraic integers. Therefore, these polynomials have coefficients in ${\mathbb Z}$; consequently, the same is true for the polynomials $f_i(x)$, as claimed. \\ Now, observe that $f_1(x)$ and $f_2(x)$ have no common roots, therefore they are relatively prime. \begin{quote} {\em Second Claim}: Let ${\mathrm{Res}}(f_1,f_2)$ be the resultant of the polynomials $f_1,f_2$. Then \begin{equation} \label{resultant f1,f2} {\mathrm{Res}}(f_1,f_2)=\pm 2^{(p-1)^2/4} p^{(p^2-1)/8}\,. \end{equation} \end{quote} Proof: We use the symbol ${\mathrm{Disc}}$ to denote the discriminant. We have \[ {\mathrm{Disc}}(f)={\mathrm{Disc}}(f_1f_2)={\mathrm{Disc}}(f_1){\mathrm{Disc}}(f_2){\mathrm{Res}}(f_1,f_2)^2\,. \] By the right-most equality in (\ref{action of sigma}) and the comments following it we see that $f_2(x)=0$ iff $f_1(-x)=0$, hence \begin{equation} \label{disc and res} {\mathrm{Disc}}(f)={\mathrm{Disc}}(f_1f_2)={\mathrm{Disc}}(f_1)^2{\mathrm{Res}}(f_1,f_2)^2\,. \end{equation} Calculation of ${\mathrm{Disc}}(f)$: \begin{align*} {\mathrm{Disc}}(f) & = p^{(2p-4)(p+1)/2}\prod_{1\leq i<j\leq p-1}\left(\omfroot{i}-\omfroot{j}\right)^2 \\ & = p^{(p+1)(p-2)}\left(\frac{-1}{p}\right)^{(p-2)(p-1)/2} \prod_{1\leq i<j\leq p-1}\left(\frac{2({\zeta}^i-{\zeta}^j)}{(1-{\zeta}^i)(1-{\zeta}^j)}\right)^2 \\ & = -2^{(p-1)(p-2)}p^{(p-2)(p+3)/2}\prod_{1\leq i<j\leq p-1}\left(\frac{{\zeta}^i-{\zeta}^j}{(1-{\zeta}^i)(1-{\zeta}^j)}\right)^2\,. \end{align*} The right-most product in the last equality is a unit times ${\lambda}^{-(p-1)(p-2)}$, therefore, $\mathrm{w}({\mathrm{Disc}}(f))=(p-1)(p-2)(p+3)/2 - (p-1)(p-2)= (p-2)(p^2-1)/2$. But, since ${\mathrm{Disc}}(f)$ is a rational integer, it follows that \begin{equation} \label{discr f} {\mathrm{Disc}}(f)=\pm 2^{(p-1)(p-2)}p^{(p-2)(p+1)/2}\,. \end{equation} Calculation of ${\mathrm{Disc}}(f_1)$: \begin{align*} {\mathrm{Disc}}(f_1) & = p^{(p-3)(p+1)/4}\prod_{0\leq i<j\leq (p-3)/2}\left(\omfroot{g^{2i}}-\omfroot{g^{2j}}\right)^2 \\ & = p^{(p+1)(p-3)/4}\left(\frac{-1}{p}\right)^{(p-3)(p-1)/8} \prod_{0\leq i<j\leq (p-3)/2}\left(\frac{2({\zeta}^{g^{2i}}-{\zeta}^{g^{2j}})}{(1-{\zeta}^{g^{2i}})(1-{\zeta}^{g^{2j}})}\right)^2 \\ & = -2^{(p-1)(p-3)/4}p^{(p-3)(p+3)/8} \prod_{0\leq i<j\leq (p-3)/2}\left(\frac{{\zeta}^{g^{2i}}-{\zeta}^{g^{2j}}}{(1-{\zeta}^{g^{2i}})(1-{\zeta}^{g^{2j}})}\right)^2\,. \end{align*} The right-most product in the last equality is a unit times ${\lambda}^{-(p-1)(p-3)/4}$, therefore, $\mathrm{w}({\mathrm{Disc}}(f_1))=(p-1)(p-3)(p+3)/8 - (p-1)(p-3)/4= (p-1)(p-3)(p+1)/8$. Since ${\mathrm{Disc}}(f_1)$ is a rational integer, it follows that \begin{equation} \label{discr f1} {\mathrm{Disc}}(f_1)=\pm 2^{(p-1)(p-3)/4}p^{(p-3)(p+1)/8}\,. \end{equation} Now the relations (\ref{disc and res}), (\ref{discr f}) and (\ref{discr f1}) imply the validity of the relation (\ref{resultant f1,f2}). \begin{quote} {\em Third Claim}: Among the integers $f_1(a),f_2(a)$ one is equal to $\pm 1$ and the other is equal to $\pm q^n$. \end{quote} Proof: By Bezout's identity, there exist polynomials $h_1(x),h_2(x)\in{\mathbb Z}[x]$ (both of degree $< (p-1)/2$) such that \[ h_1(x)f_1(x)+h_2(x)f_2(x)={\mathrm{Disc}}(f_1,f_2)= \pm 2^{(p-1)^2/4} p^{(p^2-1)/8}\,. \] We make the substitution $x \leftarrow a$ in this equality. By $f_1(a)f_2(a)=f(a)=\pm q^n$ and the fact that $(q,2p)=1$ (cf. beginning of the proof of Proposition 2.1), it follows that exactly one $f_i(a)$ is equal to $\pm 1$ and the other is equal to $\pm q^n$. \begin{quote} {\em Fourth Claim}: If $f_i(a)=\pm 1$ for some $i\in\{1,2\}$, then $a=0$. \end{quote} Proof: Let us put $f_1(x)=c_0+c_1x+\cdots +c_r x^r$, where $r=(p-1)/2$. We already know that $c_0=\pm 1$ and $c_i\in{\mathbb Z}$ for all $i$. By the very definition of the polynomial $f_1$, its roots are \[ \xi_j = \frac{1+\zeta^j}{{\omega} (1-\zeta^j)}\,,\quad j\in S\,, \] where $S$ is a complete set of quadratic residues $\bmod\,p$. \\ Since $\Phi(-1)=1$ the numerator $1+{\zeta}^j$ is a unit and it follows easily that \[ \mathrm{w}(\xi_j)= - (p+1)/2\,. \] Thus, for $k>0$, \[ \mathrm{w}(c_k) \ge -\frac{p+1}{2}\frac{p-1-2k}{2}+\frac{(p-1)(p+1)}{4} =\frac{k(p+1)}{2} \,. \] Then, for $k\geq 1$, we have $\mathrm{v}(c_k)=\mathrm{w}(c_k)/(p-1)>0$ and therefore \begin{equation} \label{p divides all c} p \mid c_k\,,\quad k=1,\ldots,(p-1)/2\,. \end{equation} Moreover, for $k\geq 2$, we have $\mathrm{v}(c_k)\geq (p+1)/(p-1)> 1$, hence $\mathrm{v}(c_k)\geq 2$ and we can write therefore \begin{equation} \label{p-adic expansion} f_1(x) \equiv c_0+c_1 x \pmod{p^2 x^2}\,. \end{equation} Next, we prove that \begin{equation} \label{val c1} \mathrm{v}(c_1)=1\,. \end{equation} First, note that $f_2(x)=f_1(-x)$, which results from the fact that the polynomial $f_1$ is of odd degree, and the polynomials $f_1$ and $f_2$ have opposite leading coefficients and roots (cf.~just before the relation (\ref{disc and res})). These observations imply that $f_2(x)=c_0-c_1x+c_2x^2-c_3x^3+\cdots $. \\ Another observation is that $c_1\neq 0$. Indeed, since $f(x)=f_1(x)f_2(x)$, the coefficient of $x^2$ in $f(x)$ is $2c_0c_2 +c_1^2$. On the other hand, by the initial definition of $f(x)$, the coefficient of $x^2$ is $-p\bincf{p}{2}$, which is odd, because $p\equiv 3\pmod 4$. Therefore, $c_1$ is odd; in particular, it is non-zero. \\ A third fact --which is a bit more than an observation-- is that \begin{equation} \label{ub for c1} |c_1| < p^2\,. \end{equation} If we prove this, then, in combination with the relation (\ref{p divides all c}) and the fact that $c_1\neq 0$, we will conclude that $\mathrm{v}(c_1)=1$. \\ Proof of (\ref{ub for c1}): Let $g_1(x)$ be, as before, the reciprocal of the polynomial $f_1(x)$. Then $c_1$ is equal, up to sign, with the sum of the roots of $g_1(x)$. But the roots of $g_1(x)$ are the reciprocals of the roots of $f_1(x)$, i.e. they are equal to $\xi_j^{-1}$, $i=1,\ldots, (p-1)/2$. Therefore (remember that $S$ is a complete set of residues $\bmod\,p$), \[ |c_1|\leq \sqrt{p}\sum_{j\in S}\left|\frac{1-{\zeta}^j}{1+{\zeta}^j}\right| =\sqrt{p}\sum_{j\in S}\left|\tan\frac{\pi j}{p}\right| = \sqrt{p}\sum_{j\in S}\frac{1}{\left|\tan\frac{\pi(p-2j)}{2p}\right|}\,. \] Since $\left|\frac{\pi(p-2j)}{2p}\right|<\frac{\pi}{2}$, it follows that $\left|\tan\frac{\pi(p-2j)}{2p}\right|>\frac{\pi}{2p}|p-2j|$, hence, \[ |c_1| <\frac{2p\sqrt{p}}{\pi}\sum_{j \in S} \frac{1}{|p-2j|}\,. \] Note that, as $j$ runs through the set $S$, the numbers $|p-2j|$ are distinct $\bmod\,p$, for, if $|p-2j_1|\equiv |p-2j_2| \pmod p$ with $j_1,j_2\in S$ and $j_1\neq j_2$, then, necessarily, $j_2 = -j_1$, which implies that $-1$ is a quadratic residue $\bmod\,p$, a contradiction. Therefore, the set $\{|p-2j|: j\in S\}$ is a subset of $\{1,\ldots,p-1\}$ with cardinality $(p-1)/2$. It is clear, therefore, that \[ \sum_{j \in S} \frac{1}{|p-2j|} \leq \sum_{k=1}^{(p-1)/2}\frac{1}{k} <\frac{3}{2}+\log\frac{p-1}{4}\,, \] from which we obtain \[ |c_1| <\frac{2p\sqrt{p}}{\pi}\left(\frac{3}{2}+\log\frac{p-1}{4}\right)\,. \] This upper bound for $|c_1|$ clearly implies $|c_1|<p^2$, as claimed. {\em Final step of the proof of Theorem \ref{th p equiv 3 mod 8}}: By our third claim above, $f_1(a)$ or $f_2(a)$ must be $\pm 1$. Since $f_2(a)=-f_1(-a)$, we may suppose that $f_1(a)=\pm 1$, i.e. $c_0+c_1a+c_2a^2+\cdots = \pm 1$. Remember that $c_0=\pm 1$, therefore, $c_0+c_1a+c_2a^2+\cdots = \pm c_0$. The $-$ sign implies $\pm 2+c_1a+c_2a^2+\cdots = 0$, clearly impossible, in view of (\ref{p divides all c}). The $+$ sign implies $0=c_1a +c_2a^2 +\cdots $. If $a\neq 0$, then, taking also into account (\ref{p-adic expansion}), we obtain $0=c_1\pmod{p^2}$ which contradicts (\ref{val c1}). This forces $a=0$ and then, by (\ref{q^n}), $q^n=\pm 2$, a contradiction. \hspace*{1mm} \hfill{$\Box$} \section{The equation $5x^2-4=y^n$} \label{section general exp eq} The third relation (\ref{shape of sols}), written as $q^n=5(20v^2-1)^2-4$, naturally leads to the study of the more general equation \begin{equation} \label{general exp eq} 5x^2 -4=y^n\,,\quad \hbox{$|x|,y>1,\,n>2$} \end{equation} in the integer unknowns $x,y,n$, where $x$ and $y$ have not, of course, the same meaning as the $x,y$ in equation (\ref{Abu eq}). First, let $n$ be even. It is well-known that the positive integer solutions of $5X^2-4=Y^2$ are given by $X=F_{2k+1}$, $Y=L_{2k+1}$ for $k>0$, where $F$ denotes the Fibonacci and $L$ the Lucas sequence; notice that $k=0$ gives $y=1$ which is excluded. Since it is known that the only Lucas number which is a pure power is $L_3=4$ (\cite{BMS}, Theorem 2), it follows that the only solution $(x,y,n)$ of the equation (\ref{general exp eq}) with even $n$ is $(2,2,4)$. From now on we suppose that $n$ is odd. \\ If $n=3$, then $(25x)^2-500=(5y)^3$. It is well known that the only integral solutions $(X,Y)$ to $Y^2=X^3+500$ are $(X,Y)=(5,\pm 25)$, corresponding to $y=1$, which has been excluded by hypothesis. Hence we may assume that ${\mathrm{gcd}}(n,3)=1$, in particular $n\ge 5$. \\ If $x$ is even then $\,5x^2-4 \equiv -4, \,16,\, 12 \pmod{32}$ implying $n\le 4$, which has already been excluded. Hence $x$ and $y$ are odd. Now we work in the field $K={\mathbb Q}({\theta})$, where ${\theta}=\sqrt{5}$. From now on and until the end of the paper we view $K$ as embedded into the real numbers with ${\theta}\mapsto \sqrt{5}=2.2360679\ldots$. The ring of integers in $K$ is ${\bf I}=\{(x+y{\theta})/2\,;\, x, y \in {\mathbb Z}\ {\rm with}\ x\equiv y \pmod{2} \}$, ${\varepsilon}=(1+\sqrt 5)/2$ is the fundamental unit. In $K$ unique factorization holds. Throughout this section, for ${\alpha}\in K$, ${\alpha}'$ will always denote the algebraic conjugate of ${\alpha}$. We factorize the equation (\ref{general exp eq}) over the field $K$ \begin{equation} \label{eq factorized over K} 5x^2-4=(x{\theta}-2)(x{\theta}+2)\,. \end{equation} If $p$ is a (rational) prime divisor of $y$, then $5x^2-4=y^n$ implies that $p$ is odd and, clearly, $5$ is a quadratic residue $\bmod\,p$. It follows that $p$ splits in $K$ and $p\equiv \pm 1 \pmod 5$. Therefore $y$ factorizes in ${\bf I}$ as $y=\pi \pi'$, where we can choose $\pi>0$ (then $\pi'$ is also positive). Notice also that $y^n\equiv 1 \pmod 5$, hence $y\equiv 1 \pmod {10}$ (remember that $y$ and $n$ are odd) and $y\ge 11$. Without loss of generality, we assume that $x$ is positive. Since $x$ is odd, $x{\theta}+2$ and $x{\theta}-2$ are coprime with $x{\theta} > y^{n/2}\geq 11^{5/2}$. Hence, there exists $k\in{\mathbb Z}$ such that $x{\theta}+2={\varepsilon}^k\pi^n$. Writing $k=\ell n+k_1$ with $-(n-1)/2\leq k_1\leq (n-1)/2$, we have $x{\theta}+2={\varepsilon}^{k_1}({\varepsilon}^{\ell}\pi)^n={\varepsilon}^{k_1}\pi_1^n$, where $\pi_1={\varepsilon}^{\ell}\pi$. The conjugate relation is $-x{\theta}+2={\varepsilon}'^{k_1}\pi_1'^n$ and summing the two relations we get \begin{equation} \label{eps and pi} {\varepsilon}^{k_1}\pi_1^n + {\varepsilon}'^{k_1}\pi_1'^n = 4\,. \end{equation} We have $\pi_1=u+v{\theta}$ or $\pi_1=(u+v{\theta})/2$, where $u,v\in{\mathbb Z}$ are unknown and in the second case $uv$ is odd. Then, for fixed $n$ and $k_1$ we obtain from (\ref{eps and pi}) \begin{equation} \label{Thue eq} T_{k_1}(u,v):= {\varepsilon}^{k_1}(u+v{\theta})^n + {\varepsilon}'^{k_1}(u-v{\theta})^n = \mbox{$4$ or $2^{n+2}$,} \end{equation} where, in the second case, $uv$ is odd. Note that the left-hand side of (\ref{Thue eq}) is a homogeneous polynomial in ${\mathbb Z}[u,v]$ of degree $n$, hence the relation (\ref{Thue eq}) implies a Thue equation. Since $T_{-k_1}(u,v)=(-1)^{k_1}T_{k_1}(u,-v)$, it suffices to consider the Thue equations $T_{k_1}(u,v)=\pm 4$ and $T_{k_1}(u,v)=\pm 2^{n+2}$ with $k_1=0,1,\ldots,(n-1)/2$, where, in the second equation, $uv$ is odd. Moreover, since the degree of the form $T_{k_1}$ is odd, we can ignore the minus sign in the right-hand sides. Using the above Thue equations we will prove that there are no solutions $(x,y,n)$ to (\ref{general exp eq}) with $n\in\{5,7,11,13\}$. Actually, we will show that for these values of $n$ the Thue equations $T_{k_1}(u,v)=2^{n+2}$ with $uv$ odd, and $T_{k_1}(u,v)=4$ are impossible for all $k_1=0,1,\ldots,(n-1)/2$. For every $n$ as above, the method is practically the same. However, as one can guess, the case $n=13$ is somewhat more complicated; so we briefly expose this case in order to illustrate how we work. Numerous Thue equations of degree $n$ arise. A practical method for the solution of such equations has been developed since long by Tzanakis and de Weger \cite{TdW1} which later was improved by Bilu and Hanrot \cite{BH} and implemented in \textsc{Pari} ($\mathtt{http://pari.math.u-bordeaux.fr/}$ and \textsc{Magma} \cite{Bosma}, \cite{magma-handbook}. We use either of these packages to solve the Thue equations that arise. We assume now that $n=13$ and we consider all $k_1$'s in $\{0,1,\ldots,6\}$.\\ $k_1=0$: Since $T_0(u,v)$ is reducible, our equations are treated by elementary means; no solutions arise. \\ $k_1=1$: Both equations $T_1(u,v)=4, 2^{15}$ are easily solved. \\ $k_1=2$: The congruences $T_2(u,v)\equiv 4, 2^{15}\pmod{13^2}$ are impossible. \\ $k_1=3$: The equation $T_3(u,v)=2^{15}$ with $uv$ odd implies solvability of the congruence $T_3(x,1)\equiv 0\pmod{2^{14}}$. But, as it is easily checked, this congruence has no solutions. The equation $T_3(u,v)=4$ remains. Since $T_3(u,v)=4u^{13}+\cdots$, we multiply by $2^{11}$ and we obtain a Thue equation $u'^{13}+65u'^{12}v+\cdots + 320000000v^{13}=2^{13}$, whose only solution is $(u',v)=(2,0)$ which we obviously reject. \\ $k_1=4$: Now, $T_4(u,v)=7u^{13}+\cdots + 234375v^{13}$. On multiplying by $7^{12}$ we obtain monic Thue equations with right-hand sides $4\cdot 7^{12}$ and $2^{15}7^{12}$. No solutions are returned. \\ $k_1=5$: Similarly to the case $k_1=2$, both congruences $T_5(u,v)\equiv 4, 2^{15}\pmod{13^2}$ are impossible. \\$k_1=6$: All coefficients of $T_6(u,v)$ are even and $\frac{1}{2}T_6(u,v)=9u^{13} +260u^{12}v+\cdots +312500v^{13}= T'_6(u,v)$, say. We thus have the Thue equations $T'_6(u,v)=2^{14}$ with $uv$ odd, and $T'_6(u,v)=2^2$. The first equation implies solvability of the congruence $T'_6(x,1)\equiv 0\pmod{2^{13}}$ with $x$ odd, which is impossible. For the second equation we are obliged to multiply by $3^{24}$ in order to obtain a monic Thue equation, as required by both \textsc{Pari} and \textsc{Magma}. The resulting equation is treated with some ``effort'' by \textsc{Magma} and no solutions are returned. On the other hand, \textsc{Pari} after several hours was still ``struggling'', so we gave up. The computational difficulties arising above, when $k_1=6$ show the limitation of the method and, indeed, for $n=17$ the computational difficulties for the solution of the resulting Thue equations, at present, seem to be insurmountable. Summing up our results so far, we have the following theorem. \begin{proposition} \label{prop small n} There are no solutions $(x,y,n)$ to the equation (\ref{general exp eq}) with $n$ divisible by at least one of the primes $3,5,7,11$ or $13$. The only solution $(x,y,n)$ to the equation (\ref{general exp eq}) with even $n$ is $(\pm 2,2,4)$. \end{proposition} {\em Computing a first upper bound for $n$}. We now fix a solution $(x,y,n)$ of the equation (\ref{general exp eq}), where, in view of Proposition \ref{prop small n}, we can assume that $n\geq 17$. Obviously, we can also assume that $n$ is prime. Based on the few observations just after the equation (\ref{eq factorized over K}), but relaxing the condition $x>0$, we see that there exists a set $\mathcal{P}$ consisting of (unordered) sets $\{\pi,\pi'\}$ such that $\pi>0$, $\pi\pi'=y$ and, if $\{\pi_1,\pi_1'\}$ and $\{\pi_2,\pi_2'\}$ are distinct elements of $\mathcal{P}$, then $\pi_2,\pi_2'$ are non-associated to both $\pi_1,\pi_1'$. \\ We modify $\mathcal{P}$ as follows: Let $\{\pi,\pi'\}\in\mathcal{P} $. There exists precisely an $m\in{\mathbb Z}$ such that ${\varepsilon}^m\leq \sqrt{{\varepsilon} y}/\pi < {\varepsilon}^{m+1}$. The last relation is equivalent to ${\varepsilon}^{2m-1}\pi^2\leq y<{\varepsilon}^{2m+1}\pi^2$. On putting ${\varepsilon}^m\pi=\pi_1$ we obtain \begin{equation} \label{ineq pi1} \frac{\pi_1}{\sqrt{{\varepsilon}}} \leq \sqrt{y} <\pi_1\sqrt{{\varepsilon}}\,,\quad \mbox{or, equivalently, $\quad \displaystyle{\sqrt{\frac{y}{{\varepsilon}}}<\pi_1\leq \sqrt{{\varepsilon} y}}$.} \end{equation} Note that $\pi_1'=(-1)^m{\varepsilon}^{-m}\pi'$, so that $\pi_1|\pi_1'|=y$. On multiplying the first relation (\ref{ineq pi1}) by $|\pi_1'|$ we get $\displaystyle{\frac{y}{\sqrt{{\varepsilon}}}\leq |\pi_1'|\sqrt{y} < y\sqrt{{\varepsilon}}}$, hence $\displaystyle{\sqrt{\frac{y}{{\varepsilon}}} \leq |\pi_1'| < \sqrt{y{\varepsilon}}}$. The last relation combined with the second relation (\ref{ineq pi1}) implies $\max\{\pi_1/|\pi_1'|\,,\,|\pi_1'|/|\pi_1|\} \leq{\varepsilon}$ and, certainly, the left-hand side of the last inequality is $>1$. We make the substitution $\pi\leftarrow \pi_1$ or $\pi\leftarrow |\pi_1'|$ according as $\pi_1/|\pi_1'|$ is $>1$ or $<1$, respectively. In this way, an ``adjusted'' set $\mathcal{P}_1$ replaces the set $\mathcal{P}$ containing elements $\{\pi,\pi'\}$ such that, \begin{equation} \label{normalized pi} \pi>0\,, \quad \pi|\pi'|=y\,,\quad 1<\pi/|\pi'|\leq {\varepsilon}\,. \end{equation} Now, in view of the relation (\ref{eq factorized over K}) and the fact that the two factors in the left-hand side are relatively prime, we must have an ideal equation $\ideal{2+x{\theta}}=\ideal{\pi}^n$ or $\ideal{2+x{\theta}}=\ideal{\pi'}^n$ for some $\{\pi,\pi'\}\in\mathcal{P}_1$, and then $\ideal{-2+x{\theta}}=\ideal{\pi'}^n$ or $\ideal{-2+x{\theta}}=\ideal{\pi}^n$, respectively. By choosing the appropriate sign for $x$ we may assume that $\ideal{2+x{\theta}}=\ideal{\pi}^n$ , from which it follows that \begin{equation} \label{eq 2+xth} 2+x{\theta}={\sigma}{\varepsilon}^k\pi^n \quad\mbox{for some $k\in{\mathbb Z}$ and ${\sigma}\in\{-1,1\}$,} \end{equation} and \begin{equation} \label{eq xth-2} x{\theta} - 2 =\frac{y^n}{x{\theta}+2}=\frac{\pi^n|\pi'|^n}{{\sigma}{\varepsilon}^k\pi^n}={\sigma}{\varepsilon}^{-k}|\pi'|^n \,. \end{equation} By (\ref{eq 2+xth}) and (\ref{eq xth-2}) we obtain \begin{equation} \label{eq power-1} {\varepsilon}^{2k}\left(\frac{\pi}{|\pi'|}\right)^n - 1 =\frac{{\sigma}{\varepsilon}^k\pi^n}{{\sigma}{\varepsilon}^{-k}|\pi'|^n} - 1 = \frac{x{\theta}+2}{x{\theta}-2}-1 =\frac{4}{x{\theta}-2}\,. \end{equation} We have $5x^2=y^n+4$, from which $|x|{\theta} >y^{n/2}$. Now we put \begin{equation} \label{def Lambda} {\Lambda} = 2k\log{\varepsilon} - n\log\frac{|\pi'|}{\pi} \,, \end{equation} so that $\displaystyle{{\Lambda}=\log({\varepsilon}^{2k}\left(\frac{\pi}{|\pi'|}\right)^n)}$ and now, by (\ref{eq power-1}), \[ |e^{{\Lambda}}-1| =\frac{4}{|x{\theta}-2|}\leq \frac{4}{|x|{\theta}-2}<\frac{4}{y^{n/2}-2}< \frac{4.0001}{y^{n/2}}\,. \] Notice that the right most side is less than $5.63\cdot 10^{-9}$ in view of the fact that $y\geq 11$ and $n\geq 17$. Therefore, \[ |{\Lambda}| < 1.01 |e^{{\Lambda}}-1| < \frac{4.0402}{y^{n/2}}\,. \] On the other hand, since the ideals $\ideal{\pi}$ and $\ideal{\pi'}$ are distinct, $\pi/|\pi'|$ is not a unit and, consequently, ${\Lambda}\neq 0$. Thus, \begin{equation} \label{ubd Lambda} 0< |{\Lambda}| < \frac{4.0402}{y^{n/2}} \end{equation} and \begin{equation} \label{ubd logLambda} \log|{\Lambda}| < -\frac{n}{2}\log y + 1.3963\,. \end{equation} Now we compare $k$ and $n$ that appear in the linear form ${\Lambda}$. Since $\log{\varepsilon}$ and $\log\frac{\pi}{\pi'}$ are both positive and, by (\ref{ubd Lambda}), $|{\Lambda}|$ is very small, it follows that $k$ is negative. By (\ref{ubd Lambda}), $|{\Lambda}|<7.6\times 10^{-6}$, therefore, in view also of (\ref{normalized pi}), \[ |k|=-k=-\frac{{\Lambda}}{2\log{\varepsilon}}+\frac{n}{2\log{\varepsilon}}\log\frac{\pi}{|\pi'|} < 7.9\times 10^{-6}+\frac{n}{2\log{\varepsilon}}\log{\varepsilon} =7.9\times 10^{-6}+\frac{n}{2}\,, \] hence \begin{equation} \label{compare k,n} |k| \leq \frac{n}{2} \,. \end{equation} Next, we consider the algebraic number $\eta:=\frac{\pi}{|\pi'|}$ appearing in ${\Lambda}$. This number is a root of the polynomial \[ (\pi X-|\pi'|)(|\pi'|X-\pi)=yX^2-(\pi^2+\pi'^2)X+y = yX^2-(a^2\pm 2y)X+y \in{\mathbb Z}\,, \] where $a=\pi+\pi'\in{\mathbb Z}$. From this we easily see that \begin{equation} \label{height eta} {\mathrm{h}}(\eta) <\frac{1}{2}(\log y +\log{\varepsilon})\,. \end{equation} Finally, we are ready to calculate a first upper bound for $n$ using Corollary 2 of \cite{Laurent}. In view of the relations (\ref{height eta}) and (\ref{normalized pi}) it is easy to estimate the quantities that are involved in that corollary. Choosing the parameter $m$ that appears in the corollary equal to 20, and taking into account that $\max(2|k|,|n|\}=n$ (cf.\ref{compare k,n})), we easily find that, if $n\geq 15100$, then \[ \log|{\Lambda}|\geq 78.8\left(\log n +\log\frac{\log y +\log{\varepsilon}+1}{\log y+\log{\varepsilon}}+0.38\right)^2(\log y+\log{\varepsilon})\,. \] This, combined with (\ref{ubd logLambda}), gives \begin{equation} \label{test corollary 2} 78.8\left(\log n +\log\frac{\log y +\log{\varepsilon}+1}{\log y+\log{\varepsilon}}+0.38\right)^2(\log y+\log{\varepsilon}) -\frac{n}{2}\log y +1.3963 > 0\,. \end{equation} Since $y\geq 11$, we easily check that the inequality (\ref{test corollary 2}) can hold only if \begin{equation} \label{first ub for n} n < 2.2\times 10^4\,. \end{equation} {\em Proving that solutions with ``small'' $y$ cannot exist}. Now we go back to our equation (\ref{general exp eq}) and we assume that $(x,y)$ is a positive solution. It is easily checked that this positiveness restriction does not prevent us from obtaining again the relations (\ref{eq power-1}) and (\ref{ubd Lambda}). We write the last inequality in the following shape: \[ \left|\frac{\log\eta}{\log{\varepsilon}}-\frac{2k}{n}\right| < \frac{4.0402}{n\log{\varepsilon}\cdot y^{n/2}}\,, \quad \eta:=\frac{\pi}{\pi'}\,. \] The right-hand side is, obviously, less than $1/(2n^2)$, which shows that $2k/n$ is a convergent to the continued fraction expansion of $\log\eta/\log{\varepsilon}$ and, moreover, the denominator of this convergent is less than $10^5$, in view of (\ref{first ub for n}). Let $a_0,a_1,a_2,\ldots$ be the partial quotients and $p_0/q_0,p_1/q_1,p_2/q_2,\ldots$ the convergents to that expansion. Let $h$ be the first subscript such that $q_h\geq 10^5$. Then, $2k/n=p_m/q_m$ for some $m\in\{0,\ldots,h-1\}$. We have now \[ \frac{1}{(a_{i+1}+2)q_i^2} < \left|\frac{\log\eta}{\log{\varepsilon}}-\frac{p_i}{q_i}\right|\,, \] hence, \[ \frac{1}{(a_{i+1}+2)n^2} < \left|\frac{\log\eta}{\log{\varepsilon}}-\frac{2k}{n}\right| <\frac{4.0402}{n\log{\varepsilon}\cdot y^{n/2}}\,, \] from which it follows that \begin{equation} \label{lb for A} 4.0402 (A+2)n >\log{\varepsilon}\cdot y^{n/2}\,,\quad A:=\max\{a_0,a_1,\ldots,a_h\}\,. \end{equation} For every $y\equiv 1\pmod{10}$ with $y<3\cdot 10^9$ and for every $\pi$ as above (there are $2^m$ such $\eta$' s, where $m$ is the number of rational prime divisors of $y$), we compute $\eta$ and the continued fraction expansion of the real number $\log\eta/\log{\varepsilon}$, and we check the validity of the relation (\ref{lb for A}). These computations can be performed with the routines of either {\sc Pari} or {\sc Magma}. We stress the fact that an ordinary precision is sufficient since the denominators of the checked convergents have at most 10 decimal digits. The whole task took around 30 hours of computations with {\sc Pari} in a usual PC; with {\sc Magma} it would take more time. It turns out that, except possibly if $n\leq 11$, this relation is {\em not} satisfied. But we already know that $n\geq 17$, hence we conclude: \begin{quote} No solutions $(x,y,n)$ to (\ref{general exp eq}) exist with $n\geq 17$ and $y\leq 3\cdot 10^9$. \end{quote} {\em Obtaining a smaller upper bound for $n$.} Now, we know that $y> 3\cdot 10^9$ (this is very important!) and we apply Theorem 2 of \cite{Laurent} to our linear form $\Lambda$. In the notation of that theorem, we choose $\rho=15.7$ and $\mu=0.57$. We obtain a new lower bound for $|\Lambda|$ which, as before, we combine with (\ref{ubd logLambda}) to obtain a complicated relation in which the only parameters that are present are $y$ and $n$. Since $y> 3\cdot 10^9$, we check that $n$ must necessarily be at most $6404$. Falling from the upper bound (\ref{first ub for n}) to $n\leq 6404$ is already a considerable improvement, but we can do much better. We turn to the main theorem of \cite{Laurent}, namely Theorem 1. Its application is somewhat complicated, as one has to choose appropriately various parameters $\rho,\mu,R_1,R_2,S_1,S_2,K,L$, but it's worth the trouble! The strategy is the following: Once we have the lower bound $y>3\cdot 10^9$ and an upper bound for $n$, we choose our parameters above in order to obtain a smaller upper bound for $n$. We repeat the process with this reduced upper bound and new parameters, and so on, as in the table below. In the table we omit the values of $K$ and $L$ because we always choose \[ L=R_1S_1\,,\quad K=\max\{\left\lceil\frac{R_2S_2-1}{R_1S_1}\right\rceil,2\}\,. \] The choice of the remaining parameters requires experiments; we cannot expose a systematic strategy for choosing them. \vspace{3mm} \tablecaption{The choice of parameters in Theorem 1 of \cite{Laurent}} \label{ParametersTable} \tablefirsthead{ \multicolumn{8}{c}{} \\[-8mm] \hline $\rho$ & $\mu$ & $R_1$ & $S_1$ & $R_2$ & $S_2$ & starting ub for $n$ & reduced ub for $n$ \\ \hline\hline } \tablehead{ \multicolumn{8}{c}{Table \ref{ParametersTable} {\small \sl(continued from previous page)}} \\ \hline $\rho$ & $\mu$ & $R_1$ & $S_1$ & $R_2$ & $S_2$ & starting ub for $n$ & reduced ub for $n$ \\ \hline\hline } \tabletail{ \hline\multicolumn{8}{c}{\small\sl continued to the next page} \\ } \tablelasttail{ } \begin{supertabular}{|l|l|c|c|c|c|r|r|} 11 & 0.6 & 1 & 20 & 400 & 110 & 6404 & 5802 \\ 11 & 0.6 & 1 & 20 & 300 & 100 & 5802 & 3956 \\ 11 & 0.6 & 1 & 20 & 211 & 80 & 3956 & 2226 \\ 10 & 0.57 & 1 & 15 & 205 & 79 & 2226 & 1949 \\ 9.95 & 0.57 & 1 & 15 & 198 & 75 & 1949 & 1783 \\ 10 & 0.555 & 2 & 8 & 195 & 75 & 1783 & 1713 \\ 10 & 0.56 & 2 & 8 & 186 & 66 & 1713 & 1452 \\ 10 & 0.555 & 2 & 7 & 184 & 63 & 1452 & 1358 \\ 10 & 0.555 & 3 & 5 & 178 & 60 & 1358 & 1251 \\ 10 & 0.555 & 4 & 4 & 173 & 60 & 1251 & 1216 \\ 11.01 & 0.6 & 4 & 4 & 172 & 49 & 1216 & 1113 \\ 10.09 & 0.59 & 4 & 3 & 170 & 48 & 1113 & 1054 \\ 11 & 0.6 & 6 & 2 & 165 & 46 & 1054 & 1003 \\ 10.9 & 0.6 & 7 & 2 & 162 & 46 & 1003 & 980 \\ 15.7 & 0.57 & 4 & 3 & 149 & 42 & 980 & 901 \\ 15.555 & 0.565 & 4 & 3 & 149 & 42 & 901 & 890 \\ 15.3 & 0.565 & 4 & 3 & 149 & 42 & 890 & 885 \\ 16.05 & 0.58 & 4 & 3 & 145 & 41 & 885 & 878 \\ 16.05 & 0.58 & 4 & 3 & 145 & 41 & 885 & 878 \\ 16.045 & 0.579 & 4 & 3 & 145 & 41 & 878 & 877 \\ 16.045 & 0.579 & 6 & 2 & 142 & 41 & 877 & 859 \\ 16.05 & 0.578 & 6 & 2 & 142 & 41 & 859 & 858 \\ 16.051 & 0.577 & 6 & 2 & 142 & 41 & 858 & 856 \\ 16.052 & 0.576 & 6 & 2 & 142 & 41 & 856 & 855 \\ 16.053 & 0.575 & 6 & 2 & 142 & 41 & 855 & 853 \\ 16.054 & 0.5744 & 6 & 2 & 142 & 41 & 853 & 852 \\ 16.054 & 0.5743 & 6 & 2 & 145 & 40 & 852 & 849 \\ 16.042 & 0.5788 & 6 & 2 & 144 & 40 & 849 & 848 \\ 16.042 & 0.5788 & 6 & 2 & 160 & 35 & 848 & 825 \\ 16.03 & 0.575 & 6 & 2 & 160 & 35 & 825 & 820 \\ \hline \end{supertabular} No further reduction of the upper bound of $n$ was possible. Summing up all the results of Section \ref{section general exp eq} we obtain the following theorem. \begin{theorem} \label{th gen exp eq} Any integer solution $(x,y,n)$ of the equation $5x^2-4=y^n$ with $y>1$ and $n>2$ must satisfy the following: (i) $n<820$. (ii) The prime divisors of $n$ are $\geq 17$ and $\leq 811$. (iii) $y>3\cdot 10^9$. \end{theorem} {\bf Remark}. In recent years, the so called ``modular approach'' to certain types of Diophantine equations --the Fermat equation being one of them-- turned out to be very succesful; see, for example, S.~Siksek's ``The modular approach to Diophantine equations'', Chapter 15 in \cite{Cohen}. Our equation $5x^2-4=y^n$ resembles the Lebesgue-Nagell equation $x^2+D=y^n$, to which the modular method applies succesfully in most cases; see \cite{BMSII}. However, as mentioned in \cite{BMSII}, the method is not succesful when $D=-a^2\pm 1$ because, in that case, there exists an obvious solution valid for every $n$. In the case of our equation we face a similar situation: the existence of the solution $(x,y)=(1,1)$ for every $n$ makes the application of the modular method ``hopeless'', according to S.~Siksek (private communication). \section{The equation $5x^2-4=y^n$ when $y$ is prime} \label{section exp eq with y prime} The main result of this section is the following \begin{theorem} \label{th with y=q} Let $q$ be an odd prime. Then, for the solutions $(x,n)$ of the equation \begin{equation} \label{eq 5x^2-4=q^n} 5x^2-4 = q^n\,,\quad x>0\,,\; n>0\,,n\neq 2 \end{equation} the following are true: (i) If $q\not\equiv 1\pmod{10}$, no solutions exist. (ii) If $q\leq 3\cdot 10^9$, no solutions with $n>2$ exist. (iii) If $(q+4)/5=\Box$, then $(x,n)=(\sqrt{\frac{q+4}{5}},1)$ is the only solution. (iv) If $(q+4)/5\neq\Box$, then at most one solution exists. (v) No solutions exist with $n>820$. (vi) No solutions exist with $n$ divisible by a prime from the set $\{2,3,5,7,11,13\}$. \end{theorem} The proof of this theorem follows from a straightforward combination of Theorem \ref{th gen exp eq}, already proved in Section \ref{section general exp eq} and Proposition \ref{at most one solution}, below. Therefore, the present section is essentially devoted to the proof of this proposition. As noted in the beginning of Section \ref{section general exp eq}, the third relation (\ref{shape of sols}), written as $5(20v^2-1)^2-4=q^n$, led us to the more general equation (\ref{general exp eq}) for which Theorem \ref{th gen exp eq} holds. In this theorem, $y$ is general and not necessarily prime as the equation $5(20v^2-1)^2-4=q^n$ would suggest. In this section, however, we will add the extra restriction that the unknown $y$ in the equation (\ref{general exp eq}) be a prime, say $y=q$, and we will prove the following theorem. \begin{proposition} \label{at most one solution} If $q$ is an odd prime, then the equation \begin{equation} \label{eq 5x^2=q^n+4} 5x^2=q^n+4\,,\quad \mbox{$x,n$ positive integers, $n$ odd} \end{equation} has at most one solution if $(q+4)/5$ is not a perfect square and exactly one solution, namely, $(x,n)=(\sqrt{(q+4)/5},1)$ if $(q+4)/5$ is a perfect square. \end{proposition} {\em Remark}: It is easy to see that the relation (\ref{eq 5x^2=q^n+4}) implies $q\equiv 1\pmod{10}$. {\bf Proof}.\: The proof of Proposition \ref{at most one solution} will be completed in three steps. \\[1mm] {\em Step 1: The gap between two solutions of (\ref{eq 5x^2=q^n+4}).} This step consists in proving that, if two solutions $(x,n)$, and $(x^{\prime},n^{\prime})$ exist, with $n^{\prime}>n$, then $n^{\prime}$ must be ``very large'' compared to $n$; see (\ref{lower bound n'}). We need first the following result. \begin{lemma} \label{2 plus x sqrt 5} Let $x$ be a positive integer and assume that \begin{equation} 2+x\sqrt{5}=\xi^{a}\,, \label{hypothesis lemma 1} \end{equation} where $\xi$ is an algebraic integer in ${\mathbb Q}(\sqrt{5})$ and $a$ is an integer $>1$. Then, $\xi=\frac{1+\sqrt{5}}{2}$, $a=3,x=1$. \end{lemma} \vspace{-5mm} {\em Proof of the lemma}. There are two possibilities for $\xi$: (I) Either $\xi=\frac{b+c\sqrt{5}}{2}$ with $b,c$ odd integers, or (II) $\xi=b+c\sqrt{5}$ with $b,c$ arbitrary integers. \\ After expansion of the right-hand side of (\ref{hypothesis lemma 1}) we obtain \[ b^{a} + 5c^{2}\binom{a}{2}b^{a-2}+\cdots=% \begin{cases} 2^{a+1} & \mbox{in case (I)}\\ 2 & \mbox{in case (II)} \end{cases}\,. \] It follows from this that, if $a$ is even, then an odd power of 2 is a square $\bmod{\,5}$ which is impossible. Therefore, $a$ is odd and \[ b^{a} + 5c^{2}\binom{a}{2}b^{a-2}+\cdots+5^{\frac{a-1}{2}}c^{a-1}\binom{a}{a-1}b = \begin{cases} 2^{a+1} & \mbox{in case (I)}\\ 2 & \mbox{in case (II)} \end{cases}\,. \] Case (II) is impossible. Indeed, note that in the left-hand side all exponents of $b$ are odd and all exponents of $c$ are even, hence $b>0$. Also, $b$ divides 2, hence $b=1$ or $2$. If $b=1$ then an obviously impossible congruence $\bmod{\,5}$ results; and if $b=2$ then $2^{a}\leq2$ which implies $a=1$, contrary to the hypothesis. \newline In case (I) we have, as before, $b>0$, $b$ is odd and $b|2^{a+1}$. Hence, $b=1$ and we have \[ 2^{a+1}=1+5\binom{a}{2}c^{2}+\cdots+5^{\frac{a-1}{2}}\binom{a}{a-1}c^{a-1} \geq2^{a+1}\,, \] where the last inequality is strict for every $c$, if $a\geq 5$ and for every $c$ with $|c|>1$ when $a=3$. Thus, to avoid the contradiction we must conclude that $a=3$ and $|c|=1$ from which it easily follows that $c=1$ and $\xi=(1+\sqrt{5})/2$. This completes the proof of Lemma \ref{2 plus x sqrt 5}. We put ${\theta} =(1+\sqrt{5})/2,{\theta} ^{\prime}=(1-\sqrt{5})/2$. These are the roots of the polynomial $x^{2}-x-1$ and ${\theta} $ is the fundamental unit of the ring of integers of ${\mathbb Q}(\theta)$. In general, for any ${{\alpha}}\in{\mathbb Q}(\theta)$ we denote by ${{\alpha}}^{\prime}$ the conjugate of ${{\alpha}}$ under the isomorphism ${\theta} \mapsto{\theta}^{\prime}$. Assume now that $(x,n)$ is a solution to equation (\ref{eq 5x^2=q^n+4}). Then $(2+x\sqrt{5})(2-x\sqrt{5})=-q^{n}$ and it is clear that the factors in the left-hand side are relatively prime as algebraic integers of ${\mathbb Q}(\theta)$. Also, every (rational) prime dividing $q$ factors into two distinct prime ideals. It follows then that there exists an algebraic integer ${{\sigma}}$ with norm $\pm q$ such that the following ideal relation is true: $(2+x\sqrt{5})=({{\sigma}})^{n}$. Then, for some $r\in {\mathbb Z}$ we have the element equation $2+x\sqrt{5}=\pm{\theta}^{r}{{\sigma}}^{n}$ and since we can assume without loss of generality that ${{\sigma}}>0$, we finally get \[ 2+x\sqrt{5}={\theta}^{r}{{\sigma}}^{n}\quad \mbox{along with the conjugate relation $2-x\sqrt{5}={\theta}'^r{\sigma}'^n$.} \] Combining the last two relations we obtain \[ 0<{{\delta}}:= \left(\frac{{\theta}^{\prime}}{{\theta}}\right)^{r}\left(\frac{{{\sigma}}^{\prime}}{{{\sigma}}}\right)^{n}+1 =\frac{4}{{\theta}^{r}{{\sigma}}^{n}}=\frac{4}{2+x\sqrt{5}}<\frac{1}{2}\,. \] Then, $\frac{1}{2}<1-{{\delta}}=-(\frac{{\theta}^{\prime}}{{\theta}})^{r}(\frac{{{\sigma}}^{\prime}}{{{\sigma}}})^{n}<1$ and in view of the inequality $|\log(1-x)|<|x|(1+|x|)$ (valid for $|x|<1/2$) we obtain \begin{equation} \label{linear form in logs 1}% \left\vert-r\log\left\vert \frac{{\theta}}{{\theta}^{\prime}}\right\vert +n\log\left\vert \frac{{{\sigma}}^{\prime}}{{{\sigma}}}\right\vert \right\vert <{{\delta}}(1+{{\delta}})\,,\quad{{\delta}}=\frac{4}{2+x\sqrt{5}} =\frac{4}{2+\sqrt{q^{n}+4}}<\frac{4}{q^{n/2}}\,. \end{equation} Now, let $(x^{\prime},n^{\prime})$ another solution to (\ref{eq 5x^2=q^n+4}) with $n^{\prime}>n$, $n^{\prime}$ odd and $x^{\prime}>0$ (hence, $x^{\prime}>x$). Exactly as before we have a relation $2+x^{\prime}\sqrt{5}={\theta}^{r^{\prime}}{{\sigma}}^{n^{\prime}}$ for a convenient $r^{\prime}\in{\mathbb Z}$ and \begin{equation} \left\vert -r^{\prime}\log\left\vert \frac{{\theta}}{{\theta}^{\prime}} \right\vert +n^{\prime}\log\left\vert \frac{{{\sigma}}^{\prime}}{{{\sigma}}} \right\vert \right\vert <{{\delta}}^{\prime}(1+{{\delta}}^{\prime})\,,\quad {{\delta}}^{\prime}=\frac{4}{2+\sqrt{q^{n^{\prime}}+4}}<\frac{4}{q^{n^{\prime}/2}} <{{\delta}}\,. \label{linear form in logs 2} \end{equation} Putting $u=\log|{\theta}/{\theta}^{\prime}|=\log((3+\sqrt{5})/2)$ and eliminating the term $\log|{{\sigma}}^{\prime}/{{\sigma}}|$ from the inequalities (\ref{linear form in logs 1}) and (\ref{linear form in logs 2}) we get $|-rn^{\prime}+r^{\prime}n|u<n^{\prime}{{\delta}}(1+{{\delta}})+n{{\delta}}^{\prime}(1+{{\delta}}^{\prime}) <2n^{\prime}{{\delta}}(1+{{\delta}})$, i.e. \begin{equation} |-rn^{\prime}+r^{\prime}n|<\frac{2n^{\prime}}{u}{{\delta}}(1+{{\delta}})\,. \label{linear form r and r'}% \end{equation} The left-hand side in (\ref{linear form r and r'}) is non-zero. Indeed, in the opposite case we would have $\frac{r}{n}=\frac{r^{\prime}}{n^{\prime}}=\,\mbox{(say)\,}\frac{r_{1}}{n_{1}}$ with $(r_{1},n_{1})=1$. Then, $r=ar_{1},n=an_{1}$, $r^{\prime}=br_{1},n^{\prime}=bn_{1}$ for some positive odd integers $a,b$ with $a<b$ and, moreover, $2+x^{\prime}\sqrt{5}=({\theta}^{r_{1}}{{\sigma}}^{n_{1}})^{b}$. By Lemma \ref{2 plus x sqrt 5} we conclude that $x^{\prime}=1$, contrary to the fact that $x^{\prime}>x>1$. \\ We conclude therefore that the left-hand side of (\ref{linear form r and r'}) is $\geq1$, from which it follows that \begin{equation} n^{\prime}>\frac{u}{2}{{\delta}}^{-1}(1+{{\delta}})^{-1}, \label{lower bound n'} \end{equation} which shows that, the larger solution $n^{\prime}$ is \textquotedblleft far away\textquotedblright\ from the smaller solution $n$; specifically, it is of the size of $q^{n/2}$. This fact will play an important role below. \\[1mm] {\em Step 2: Application of Hypergeometric Polynomials.} At this second step we adapt to our equation the method of F.~Beukers in \cite{Beu1} and \cite{Beu2}. As a result we prove Lemma \ref{technical lemma} below, after which the final step for the proof of Proposition \ref{at most one solution} is not difficult. In that method one uses as a tool the hypergeometric polynomials, the properties of which we remind immediately below. \\ Given the real numbers ${{\alpha}},{{\beta}},{\gamma}$ where ${\gamma}$ is not zero or a negative integer, we define the \emph{hypergeometric function} (with parameters ${{\alpha}},{{\beta}},{\gamma}$) \[ F({{\alpha}},{{\beta}},{\gamma},z)=1+\frac{{{\alpha}}{{\beta}}}{{\gamma}}z +\sum_{k=2}^{\infty}\frac{1}{k!}\frac{{{\alpha}}({{\alpha}}+1) \cdots({{\alpha}}+k-1){{\beta}}({{\beta}}+1)\cdots({{\beta}}+k-1)}{{\gamma}({\gamma}+1) \cdots({\gamma}+k-1)}z^{k} \] which converges for every complex number $z$ with $|z|<1$ and, in case that ${\gamma}>{{\alpha}}+{{\beta}}$, it also converges for $z=1$. Let $n_{2}>n_{1}>0$ be integers. Put $n=n_{1}+n_{2}$ and define \[ G(z)=F(-n_{2}-1/2,-n_{1},-n,z)\,,\quad H(z)=F(-n_{1}+1/2,-n_{2},-n,z)\,. \] By the definition of $G$ it is easy to see that \[ G(z)=\sum_{k=0}^{n_{1}}\binom{n_{2}+1/2}{k}\binom{n_{1}}{k}\binom{n}{k}^{-1}(-z)^{k}\,, \] which, in particular, shows that, for any real number $z<0$, $G(z)$ is positive. We will use the following properties: \vspace{-5mm} \begin{enumerate} \item \label{prop 1} $G(z)$ and $H(z)$ are polynomials in $z$ of degrees $n_{1} $ and $n_{2}$, respectively. Moreover, the polynomials $\binom{n}{n_{1}}G(4z) $ and $\binom{n}{n_{1}}H(4z)$ have integer coefficients. \item \label{prop 2} $|G(z)-(1-z)^{1/2}H(z)|<G(1)|z|^{n+1}$ for $|z|<1$. \item \label{prop 3} $G(1)<G(z)<G(0)<1$ for $0<z<1$. \item \label{prop 4} If $G^{*}(z)$ is the polynomial resulting from $G(z)$ when $n_{1},n_{2}$ are respectively replaced by $n_{1}+1,n_{2}+1$ and $H^{*}(z)$ is defined analogously, then \[ G^{*}(z)H(z)-G(z)H^{*}(z)=cz^{n+1} \] for some non-zero constant $c$. \item \label{prop 5} $\displaystyle{|G(z)| < \left( 1+\frac{|z|}{2}\right)^{n_{2}+1}} $ for any $z$. \end{enumerate} \vspace{-4mm} For the proof of the first four properties see Lemmas 1,2,3 and 4 in \cite{Beu1}. For the proof of the fifth property see relation (1.10), page 226 of \cite{TzW}. Now we are in a position to prove the main result of this step. \begin{lemma} \label{technical lemma} Let $(x,n)$ be a solution to (\ref{eq 5x^2=q^n+4}), where, as always, $x>0$ and $n\geq1$ is odd; we assume, moreover, that $q^{n}>600$. Let $r,s$ be positive integers such that $q^{n}\geq2^{6+4s/r}$ and define the positive real number $\nu$ by means of the relation \[ q^{n\nu}=2.007\times(4.03q^{n})^{r/s}\,. \] Finally, let $N=q^{n^{\prime}}$ where $n^{\prime}>n$ and let $y$ be any integer. Then, \[ \left\vert \frac{y\sqrt{5}}{N^{1/2}}-1\right\vert >\frac{0.27}{q^{n(3+\nu/2)}}q^{n/s}N^{-(1+\nu)/2}\,. \] \end{lemma} {\em Proof of the lemma}. Let $n_{2}>n_{1}$ be positive integers which will be specified later and $m=n_{1}+n_{2}$. Put $z=-q^{-n}$. Then, $|4z|<1$ so that $G(4z)$ and $H(4z)$ are meaningful. By properties \ref{prop 2} and \ref{prop 3} of the polynomials $G$ and $H$ we have \[ \left\vert G(4z)-H(4z)(1+\frac{4}{q^{n}})^{1/2}\right\vert <G(1)\left(\frac{4}{q^{n}}\right)^{m+1} <\left(\frac{4}{q^{n}}\right)^{m+1}\,, \] hence \begin{equation} \label{inequality with G,H}% \left\vert \binom{m}{n_{1}}G(4z)-\binom{m}{n_{1}}H(4z)\frac{x\sqrt{5}}{q^{n/2}}\right\vert <\binom{m}{n_{1}}\left(\frac{4}{q^{n}}\right)^{m+1}\,. \end{equation} By property \ref{prop 1} and the fact that $G(x)>0$ for any negative real number $x$, \[ \binom{m}{n_{1}}G(4z)=\frac{A}{q^{nn_{1}}}\quad\mbox{for some positive $A\in{\mathbb Z}$} \] and similarly, \[ \binom{m}{n_{1}}H(4z)=\frac{B}{q^{nn_{2}}}\quad\mbox{for some $B\in{\mathbb Z}$.} \] Then, (\ref{inequality with G,H}) implies $\left\vert \dfrac{A}{q^{nn_{1}}}-\dfrac{Bx}{q^{nn_{2}}}\dfrac{\sqrt{5}}{q^{n/2}}\right\vert < \dbinom{m}{n_{1}}\left(\dfrac{4}{q^{n}}\right)^{m+1}$, from which \[ \left\vert 1-\frac{Bx\sqrt{5}}{Aq^{n(n_{2}-n_{1}+1/2)}}\right\vert <2^{m-1}\frac{q^{nn_{1}}}{A}\frac{2^{2(m+1)}}{q^{n(m+1)}} =\frac{2^{3m+1}}{Aq^{n(n_{2}+1)}}\,. \] Now, let us put ${{\epsilon}}=\left\vert \frac{y\sqrt{5}}{N^{1/2}}-1\right\vert$, so that, from the above inequality we have \begin{equation} \left\vert \frac{y}{N^{1/2}}-\frac{Bx}{Aq^{n(n_{2}-n_{1}+1/2)}}\right\vert <\frac{1}{\sqrt{5}}\left( {{\epsilon}}+\frac{2^{3m+1}}{Aq^{n(n_{2}+1)}}\right)\,. \label{estimate difference}% \end{equation} Let ${\lambda}=\lceil\frac{n^{\prime}-n}{2n}\rceil$. Then, \begin{equation} q^{n({\lambda}-1)}<\left( \frac{N}{q^{n}}\right)^{1/2}\leq q^{n{\lambda}}\,. \label{define la}% \end{equation} Now comes the moment to choose $n_{1},n_{2}$. First we choose $n_{1}$ to satisfy \begin{equation} \frac{r}{s}{\lambda}\leq n_{1}\leq\frac{r}{s}{\lambda}+\frac{2s-1}{s}\,. \label{choose n1}% \end{equation} We must keep in mind that there are exactly two consecutive positive integers in the interval $[r{\lambda}/s\,,\,(r{\lambda}+2s-1)/s]$; this is a simple exercise. Choose now $n_{2}$ by setting $n_{2}=n_{1}+{\lambda}>n_{1}$ and remember that $m=n_{1}+n_{2}=2n_{1}+{\lambda}$. Moreover, we will need below that the left-hand side of (\ref{estimate difference}) be non-zero. In the next lines we show that we can choose $n_{1}$ in such a way that this requirement be satisfied. \newline Suppose that for the smaller integer $n_{1}$ in the interval $[r{\lambda}/s\,,\,r{\lambda}+2s-1)/s]$ the left-hand side of (\ref{estimate difference}) is zero. Then, we can repeat the above process with $n_{1}^{\prime}:=n_{1}+1$ in place of $n_{1}$ ($n_{1}^{\prime}$ still belongs to this interval), $n_{2}^{\prime}:=n_{2}+1$ in place of $n_{2}$ and $m^{\prime}:=n_{1}^{\prime}+n_{2}^{\prime}$ in place of $m$, so that the polynomials $G$ and $H$ will be replaced by $G^{\ast}$ and $H^{\ast}$ respectively, and the integers $A,B$ by some other integers, say, $A^{\ast},B^{\ast}$. Then, we will obtain an inequality analogous to (\ref{estimate difference}), namely, \[ \left\vert \frac{y}{N^{1/2}}-\frac{B^{\ast}x}{A^{\ast}q^{n(n_{2}^{\prime}-n_{1}^{\prime}+1/2)}}\right\vert <\frac{1}{\sqrt{5}}\left({{\epsilon}}+\frac{2^{3m^{\prime}+1}}{A^{\ast}q^{n(n_{2}^{\prime}+1)}}\right)\,. \] If the left-hand side were again zero, then we would have $B/A=B^{\ast}/A^{\ast}$ (note that $n_{2}^{\prime}-n_{1}^{\prime}=n_{2}-n_{1}$), which would easily imply that $z=-4/q^{n}$ is a zero of the function $G^{\ast}\cdot H-G\cdot H^{\ast}$ and this contradicts property \ref{prop 4} of the polynomials $G,H$. We conclude therefore that for at least one integer $n_{1}$ satisfying (\ref{choose n1}), the left-hand side of (\ref{estimate difference}) is non-zero and from now on we assume that we have selected such an $n_{1}$. We now rewrite the term $Aq^{n(n_{2}-n_{1}+1/2)}$ appearing in the left-hand side of (\ref{estimate difference}). We first observe that (\ref{define la}) implies $q^{n^{\prime}/2}\leq q^{n({\lambda}+1/2)}$ which shows that $q^{n(2{\lambda}+1)}=q^{n^{\prime}}q^{2\mu}$ for some non-negative integer $\mu$. Consequently, on putting $q^{\mu}=A_{0}$ (a positive integer), we have \[ Aq^{n(n_{2}-n_{1}+1/2)}=Aq^{n({\lambda}+1/2)}=Aq^{n^{\prime}/2}A_{0} =A_{0}AN^{1/2}\,. \] Going back to (\ref{estimate difference}), we get \begin{align*} \frac{1}{\sqrt{5}}\left( {{\epsilon}}+\frac{2^{3m+1}}{Aq^{n(n_{2}+1)}}\right) & >\left\vert \frac{y}{N^{1/2}}-\frac{Bx}{Aq^{n(n_{2}-n_{1}+1/2)}}\right\vert =\left\vert \frac{y}{N^{1/2}}-\frac{Bx}{A_{0}AN^{1/2}}\right\vert =\frac{|A_{0}Ay-Bx|}{A_{0}|A|N^{1/2}}\\ & \geq\frac{1}{A_{0}|A|N^{1/2}}=\frac{1}{|A|q^{n({\lambda}+1/2)}}\,, \end{align*} from which \begin{equation} {{\epsilon}}|A|q^{n({\lambda}+1/2)}+2^{3m+1}q^{-n(n_{1}+1/2)}>\sqrt{5}\,. \label{sum larger than 1}% \end{equation} We estimate separately the second summand in the left-hand side of (\ref{sum larger than 1}). By the hypothesis on the lower bound of $q^{n}$ and (\ref{define la}) we have \[ \frac{2^{3m+1}}{q^{nn_{1}}}<\frac{2^{6n_{1}+3{\lambda}+1}}{q^{nn_{1}}} \leq\frac{2^{6n_{1}+3{\lambda}+1}}{2^{(6+4s/r)n_{1}}}=2^{3{\lambda}+1-4sn_{1}/r} \leq2^{3{\lambda}+1-4{\lambda}}=2^{1-{\lambda}}\leq 1\,, \] which shows that the second summand in the left-hand side of (\ref{sum larger than 1}) is $\leq q^{-n/2}<600^{-1/2}$. This shows that the first summand in the left-hand side of (\ref{sum larger than 1}) is larger than $5^{1/2}-600^{-1/2}>2.195$. Then, remembering also how $A$ has been defined and using property \ref{prop 5} of the polynomial $G$, we get \begin{align*} 2.195 & \leq{{\epsilon}}|A|q^{n({\lambda}+1/2)}={{\epsilon}}q^{n(n_{1}+{\lambda}+1/2)}\binom{m}{n_{1}}|G(-4/q^{n})| \\ & <{{\epsilon}}q^{n(n_{1}+{\lambda}+1/2)}\cdot 2^{m-1}\left(1+\frac{2}{q^{n}}\right)^{n_{2}+1} \\ & \leq{{\epsilon}}q^{n(n_{1}+{\lambda}+1/2)}\cdot 2^{m-1}\left(1+\frac{2}{q^{n}}\right)^{m} ={{\epsilon}}q^{n/2}q^{n(n_{1}+{\lambda})}\cdot\frac{1}{2}\left(2+\frac{4}{q^{n}}\right)^{m} \\ & <\frac{{{\epsilon}}}{2}q^{n/2}q^{n(n_{1}+{\lambda})}\times2.007^{m} \quad\mbox{(since $q^n>600$)} \\ & =\frac{{{\epsilon}}}{2}q^{n/2}q^{n{\lambda}(1+n_{1}/{\lambda})}\times2.007^{{\lambda}(1+2n_{1}/{\lambda})} \\ & \leq\frac{{{\epsilon}}}{2}q^{n/2}q^{n{\lambda}(\frac{r}{s} +\frac{2s-1}{{\lambda}s}+1)}\times2.007^{{\lambda}(\frac{2r}{s}+\frac{4s-2}{{\lambda}s}+1)} \\ & =\frac{{{\epsilon}}}{2}q^{n/2}q^{n(2s-1)/s}\times2.007^{(4s-2)/s}\cdot(q^{n(1+r/s)}\times2.007^{1+2r/s})^{{\lambda}}\\ & <\frac{{{\epsilon}}}{2}q^{n/2}q^{n(2s-1)/s}\times2.007^{(4s-2)/s}\cdot(2.007\cdot(4.03q^{n})^{r/s}q^{n})^{{\lambda}}\\ & =\frac{{{\epsilon}}}{2}q^{n/2}q^{n(2s-1)/s}\times2.007^{(4s-2)/s}\cdot q^{n(1+\nu){\lambda}} \quad\mbox{(by the definition of $\nu$)}\\ & <\frac{{{\epsilon}}}{2}q^{n/2}q^{n(2s-1)/s}\times2.007^{4}\cdot q^{n(1+\nu){\lambda}}\,, \end{align*} from which we immediately get \[ 0.27q^{n(-\frac{5}{2}+\frac{1}{s})}<{{\epsilon}}q^{n{\lambda}(1+\nu)}<{{\epsilon}}(Nq^{n})^{(\nu+1)/2} \] (the right-most inequality being true because $q^{n{\lambda}}<(Nq^{n})^{1/2}$ in view of (\ref{define la})), and hence the claimed lower bound for ${{\epsilon}}=|\frac{y\sqrt{5}}{N^{1/2}}-1|$. This completes the proof of Lemma \ref{technical lemma}. \\[1mm] {\em Step 3: Completion of the proof of Theorem \ref{at most one solution}}. Assume that, if $(q+4)/5$ is not a perfect square there exists a solution to equation (\ref{eq 5x^2=q^n+4}) and if $(q+4)/5$ is a perfect square there exists a solution to this equation besides the obvious one which results from the relation $5(\sqrt{(q+4)/5})^2=q+4$. Thus, in both cases, our assumptions in particular imply that there exists a solution $(x_0,n_0)$ with $n_0>1$, hence, by Theorem \ref{th gen exp eq}, we must have $q>3\cdot 10^9$. Let $(x,n)$ be the least solution to equation (\ref{eq 5x^2=q^n+4}). In order to prove the theorem, we will assume that a larger solution $(x',n')$ to (\ref{eq 5x^2=q^n+4}) exists and we will arrive at a contradiction. \\ We put $N=q^{n^{\prime}}$, so that $N+4=5x^{\prime2}$, from which we get \[ \frac{4}{N}=\left( \frac{\sqrt{5}x^{\prime}}{N^{1/2}}-1\right) \left(\frac{\sqrt{5}x^{\prime}}{N^{1/2}}+1\right) =\left( \frac{\sqrt{5}x^{\prime}}{N^{1/2}}-1\right) \left( \frac{\sqrt{N+4}}{\sqrt{N}}+1\right) > 2\left(\frac{\sqrt{5}x^{\prime}}{N^{1/2}}-1\right)\,, \] therefore \[ 0<\frac{\sqrt{5}x^{\prime}}{N^{1/2}}-1<\frac{2}{N}\,. \] We apply Lemma \ref{technical lemma} with $y=x^{\prime}$, $r=1,s=2$; then it is easy to check that $\nu<0.7178$ and by the conclusion of the lemma and the last displayed inequality we get \[ 2N^{-1}>\frac{\sqrt{5}x^{\prime}}{N^{1/2}}-1>0.27\times q^{-n(5+\nu)/2}N^{-(1+\nu)/2}\,, \] hence \[ \frac{(1-\nu)}{2}m^{\prime}\log q<\log(7.408)+\frac{5+\nu}{2}n\log q<\frac{5.627+\nu}{2}n\log q\,, \] from which \[ n^{\prime}<\frac{5.627+\nu}{1-\nu}n\,. \label{upper bound n'} \] On the other hand, recalling that $u=\log((3+\sqrt{5})/2)$ and ${{\delta}=}\frac{4}{2+\sqrt{q^{n}+4}}$, we have in view of (\ref{lower bound n'}), \begin{align*} n^{\prime} & >\frac{u}{2}{{\delta}}^{-1}(1+{{\delta}})^{-1} =\frac{u}{2}\cdot\frac{2+\sqrt{q^{n}+4}}{4}\left(1+\frac{4}{2+\sqrt{q^{n}+4}}\right)^{-1} \\ & =\frac{u}{8}\cdot\frac{(2+\sqrt{q^{n}+4})^{2}}{6+\sqrt{q^{n}+4}} > \frac{u}{8}\cdot\frac{(2+q^{n/2})^{2}}{6+q^{n/2}} >0.12\times\frac{(2+q^{n/2})^{2}}{6+q^{n/2}}\,. \end{align*} Combining this lower bound for $n^{\prime}$ with (\ref{upper bound n'}), we get the following relation: \[ 0.12\times\frac{(2+q^{n/2})^{2}}{6+q^{n/2}}<\frac{5.627+\nu}{1-\nu}n\,. \] By the definition of $\nu$, $(q^{n})^{\nu}=2.007\times(4.03q^{n})^{1/2}$. Solving for $\nu$ and substituting into the above inequality we obtain \begin{equation} \label{impossible ineq}% n\frac{6.127n\log q+{\gamma}}{0.5n\log q-{\gamma}}>0.12\frac{(2+q^{n/2})^2}{6+q^{n/2}}\,,\quad {\gamma}=\log (2.007\cdot 4.03^{0.5})\,. \end{equation} However, in view of the large size of $q$ we easily check that (\ref{impossible ineq}) is {\em not} satisfied and this contradiction proves that the solution $(x^{\prime},n^{\prime})$ cannot exist. \hspace*{1mm} \hfill{$\Box$}
\section{Introduction} Let ${\mathbb F}_q$ be the finite field of order $q$, where $q=p^e$, $p$ is an odd prime, and $e$ is a positive integer. Let $PG(2,q)$ denote the classical projective plane. That is, the points and lines of $PG(2,q)$ are the $1$-dimensional subspaces and $2$-dimensional subspaces of ${\mathbb F}_q^3$, respectively, and the incidence is the natural inclusion. An {\it oval} in $PG(2,q)$ is a set of $q+1$ points, no three of which are collinear. A {\it conic} in $PG(2,q)$ is the set of points $\langle (x,y,z)\rangle$ satisfying a nonzero quadratic form. A conic is said to be {\it nondegenrate} if it does not contain an entire line of $PG(2,q)$. It is well known \cite{hirsch} that every nondegenerate conic is an oval; moreover, by a linear change of coordinates, any nondegenerate conic is equivalent to $$\mathcal{O} = \{\langle(1,t,t^2)\rangle\mid t\in {\mathbb F}_q\}\cup\{\langle(0,0,1)\rangle\},$$ the set of projective solutions of the nondegenerate quadratic form $$Q(X_0,X_1,X_2) = X_1^2-X_0 X_2$$ over ${\mathbb F}_q$. In the case where $p$ is odd, Segre \cite{seg} proved that an oval in $PG(2,q)$ must be a nondegenerate conic. Therefore in $PG(2,q)$, $q$ an odd prime power, ovals and nondegenerate conics are essentially the same objects. In the rest of this paper, we will always assume that $p$ is an odd prime, and use the above $\mathcal{O}$ as our ``standard" conic. With respect to $\mathcal{O}$, a line $\ell$ is called {\it secant}, {\it tangent}, or {\it skew} according as $|\ell\cap\mathcal{O}|=2$, $1$, or $0$. We use $Se$, $T$, and $Sk$ to denote the set of secant lines, the set of tangent lines, and the set of skew lines, respectively. Since $\mathcal{O}$ is an oval, every line of $PG(2,q)$ must fall into one of these sets. A point of $PG(2,q)$ is called {\it external}, {\it absolute}, or {\it internal} with respect to $\mathcal{O}$ according as it lies on 2, 1, or 0 tangent lines to $\mathcal{O}$. We use $E$ and $I$ to denote the set of external points and the set internal points, respectively. Let ${\mathbf A}$ be the line-point incidence matrix of $PG(2,q)$. That is, the rows of ${\mathbf A}$ are labeled by the lines of $PG(2,q)$, the columns of ${\mathbf A}$ are labeled by the points of $PG(2,q)$, and the $(\ell, P)$- entry ${\mathbf A}$ is 1 if $P\in \ell$, $0$ otherwise. The following result is a special case of the well-known $p$-rank formula for the incidence matrix of the Singer design (see \cite{smith}, \cite{MacMann}, and \cite{GDel}). \begin{Theorem}\label{rankofA} rank$_p({\mathbf A})$ = $\binom{p+1}{2}^e+1$. \end{Theorem} In a recent paper \cite{DMM}, Droms, Mellinger and Meyer considered the following partition of ${\mathbf A}$ into submatrices: \begin{displaymath} {\mathbf A}=\left(\begin{array}{ccc} {\mathbf A}_{11} & {\mathbf A}_{12} & {\mathbf A}_{13} \\ {\mathbf A}_{21} & {\mathbf A}_{22} & {\mathbf A}_{23} \\ {\mathbf A}_{31} & {\mathbf A}_{32} & {\mathbf A}_{33} \\ \end{array}\right), \end{displaymath} where the rows of ${\mathbf A}_{11}, {\mathbf A}_{21}$ and ${\mathbf A}_{31}$ are labeled by the tangent, skew, and secant lines, respectively, and the columns of ${\mathbf A}_{11}, {\mathbf A}_{12}$ and ${\mathbf A}_{13}$ are labeled by the absolute, internal, and external points, respectively. These authors used the submatrices $A_{ij}$ to construct binary linear codes. Some of these codes are good examples of structured low-density parity-check (LDPC) codes, cf \cite{DMM}. Based on computational evidence, the authors of \cite{DMM} made conjectures on the dimensions of the binary LDPC codes. These conjectures were investigated in \cite{thesis}, and will be proved in forthcoming papers. In this paper, we are interested in the $p$-ranks of the submatrices ${\mathbf A}_{ij}$, where $p$ is the characteristic of ${\mathbb F}_q$. We mention that several authors have considered the $p$-ranks of various submatrices of ${\mathbf A}$. Let ${\mathbf A}_{nonsec}$ (resp. ${\mathbf A}_{sec}$) denote the submatrices of ${\mathbf A}$ obtained by deleting the rows of ${\mathbf A}$ indexed by the secant (resp. tangent and skew) lines. That is, \begin{displaymath} {\mathbf A}_{nonsec} =\left(\begin{array}{ccc} {\mathbf A}_{11} & {\mathbf A}_{12} & {\mathbf A}_{13} \\ {\mathbf A}_{21} & {\mathbf A}_{22} & {\mathbf A}_{23} \\ \end{array}\right)\;\text{and}\; {\mathbf A}_{sec}=\left(\begin{array}{ccc} {\mathbf A}_{31} & {\mathbf A}_{32} & {\mathbf A}_{33}\\ \end{array}\right). \end{displaymath} Then Blokhuis and Moorhouse \cite{blokhuis} proved the following theorem: \begin{Theorem}\label{sub_1} Use the above notation, we have \begin{itemize} \item[(i)] $rank_p {\mathbf A}_{nonsec}$ = $\binom{p+1}{2}^e+1$, \item[(ii)] $rank_p {\mathbf A}_{sec}$ = $\binom{p+1}{2}^e$. \end{itemize} \end{Theorem} The authors of \cite{blokhuis} also computed the 2-ranks of ${\mathbf A}_{nonsec}$ and ${\mathbf A}_{sec}$ when $q$ is a power of 2. Carpenter \cite{carp}, and later Kamiya and Fossorier \cite{kf} computed the 2-ranks of some more refined submatrices of ${\mathbf A}$ when $q$ is a power of 2. Our main result in this paper is the following theorem: \begin{Theorem}\label{you} Let ${\mathbf A}_{ij}$, $1\le i, j \le 3$, be the submatrices of ${\mathbf A}$ defined above. Then \begin{itemize} \item[(i)] rank$_p({\mathbf A}_{11})$ = $q+1$, \item[(ii)] rank$_p({\mathbf A}_{12})$ = rank$_p({\mathbf A}_{21})$ = $0$, \item[(iii)] rank$_p({\mathbf A}_{13})$ = rank$_p({\mathbf A}_{31})$ = $q$, \item[(iv)] rank$_p({\mathbf A}_{22}) = $rank$_p({\mathbf A}_{23})$ = rank$_p({\mathbf A}_{32})$ = $\binom{p+1}{2}^e - q$, \item[(v)] rank$_p({\mathbf A}_{33})$ = $\binom{p+1}{2}^e$. \end{itemize} \end{Theorem} The paper is organized as follows. In Section 2, we collect basic geometric results related to a conic in $PG(2,q)$. In Section 3, we first construct several vector spaces of polynomials over ${\mathbb F}_q$; we then convert the problem of computing the $p$-ranks of the above submatrices into the one of calculating the dimensions of these polynomial spaces over ${\mathbb F}_q$. Finally we prove Theorem~\ref{you} by explicitly computing the dimensions of the polynomial spaces constructed. The proofs mainly rely on the following Nullstellensatz proved in \cite{blokhuis}: \begin{Lemma}\label{new} Let $F_d[X_0,X_1,X_2]$ be the vector space of homogeneous polynomials of degree $d$ in ${\mathbb F}_q[X_0,X_1,X_2]$, together with 0. Let $f(X_0,X_1,X_2)\in F_d[X_0,X_1,X_2]$, with $d\leq q-1$ and $q$ odd. Define $Q(X_0,X_1,X_2)=X_1^2-X_0X_2$. \begin{itemize} \item[(i)] If $f(x_0,x_1,x_2) = 0$ whenever $Q(x_0,x_1,x_2)$ is a non-zero square, where $(x_0,x_1,x_2)\in {\mathbb F}_q^3$, then $f = 0$. \item[(ii)] If $f(x_0,x_1,x_2) = 0$ whenever $Q(x_0,x_1,x_2)$ is a non-square or zero, where $(x_0,x_1,x_2)\in {\mathbb F}_q^3$, then $f=0$. \end{itemize} \end{Lemma} \section{Some known geometric results related to a conic in $PG(2,q)$} A {\it correlation} of $PG(2,q)$ is a bijection sending points to lines and lines to points that reverses inclusion. A {\it polarity} of $PG(2,q)$ is a correlation of order $2$. The image of a point ${\mathbf P}$ under a correlation $\sigma$ is denoted by ${\mathbf P}^\sigma$, and that of a line $\ell$ is denoted by $\ell^\sigma$. It can be shown \cite[p.~181]{hirsch} that the nondegenerate quadratic form $Q(X_0,X_1,X_2)$ = $X_1^2-X_0X_2$ induces a polarity $\sigma$ of $PG(2,q)$, which can be represented by the matrix \begin{displaymath} M=\left(\begin{array}{ccc} 0 & 0 & -\frac{1}{2} \\ 0 & 1 & 0 \\ -\frac{1}{2} & 0 & 0 \\ \end{array}\right). \end{displaymath} For example, if ${\mathbf P}=\langle(x,y,z)\rangle$ is a point of $PG(2,q)$, then its image under $\sigma$ is ${\mathbf P}^{\sigma}=\langle M(x,y,z)^{\top}\rangle=\langle(z,-2y,x)^{\top}\rangle$. \begin{Lemma}\label{bijection}$(${\rm{\cite[p.~181]{hirsch}}}$)$ The polarity $\sigma$ defines the following three bijections: $$\sigma:\;I\rightarrow\;Sk,$$ $$\sigma:\;E\rightarrow\;Se,$$ and $$\sigma:\;\mathcal{O}\rightarrow\;T.$$ \end{Lemma} For convenience, we denote the set of all non-zero squares of ${\mathbb F}_q$ by $\;\Box_q$, and the set of non-squares by $\;\not\!\!\Box_q$. Then we have the following lemma. \begin{Lemma}\label{lem1}$(${\rm{\cite[p.~182]{hirsch}}}$)$ A line $\langle(r,m,n)^{\top}\rangle$ of $PG(2,q)$ is skew, tangent, or secant to $\mathcal{O}$ if and only if $m^2-4nr \in \;\not\!\!\Box_q$, $m^2-4nr = 0$, or $m^2-4nr\in \;\Box_q$, respectively. \end{Lemma} Using the polarity $\sigma$ induced by $Q$ and the previous lemma, we have: \begin{Lemma}\label{lem5}$(${\rm{\cite[p.~182]{hirsch}}}$)$ A point ${\mathbf P}=\langle(x,y,z)\rangle$ of $PG(2,q)$ is internal, external, or absolute if and only if $y^2-xz \in \;\not\!\!\Box_q$, $y^2-xz \in \;\Box_q$, or $y^2-xz = 0$, respectively. \end{Lemma} \begin{Lemma}{\rm (\cite[p.~170]{hirsch})} We have the following tables: \begin{table}[htp] \begin{center} \caption{Number of lines of various types} \begin{tabular}{ccc} \hline {Tangent lines} & {Skew lines} & {Secant lines} \\ \hline {$q+1$} & $\frac{q(q-1)}{2}$ & $\frac{q(q+1)}{2}$ \\ \hline \end{tabular} \label{table:points} \end{center} \end{table} \begin{table}[htp] \begin{center} \caption{Number of points of various types} \begin{tabular}{ccc} \hline {Absolute points} & {Internal points} & {External points}\\ \hline $q+1$& $\frac{q(q-1)}{2}$ & $\frac{q(q+1)}{2}$ \\ \hline \end{tabular} \label{table:lines} \end{center} \end{table} \begin{table}[htp] \begin{center} \caption{Number of points on lines of various types} \bigskip \begin{tabular}{cccc} \hline {Name} & {Absolute points} & {External points} & {Internal points} \\ \hline {Tangent lines} & $1$ & $q$ & $0$ \\ {Secant lines} & $2$ & $\frac{q-1}{2}$ & $\frac{q-1}{2}$ \\ {Skew lines} & $0$ & $\frac{q+1}{2}$ & $\frac{q+1}{2}$\\ \hline \end{tabular} \label{table:tab1} \end{center} \end{table} \begin{table}[htp] \begin{center} \caption{Number of lines through points of various types} \bigskip \begin{tabular}{cccc} \hline {Name} & {Tangent lines} & {Secant lines} & {Skew lines} \\ \hline {Absolute points} & $1$ & $q$ & $0$ \\ {External points} & $2$ & $\frac{q-1}{2}$ & $\frac{q-1}{2}$ \\ {Internal points} & $0$ & $\frac{q+1}{2}$ & $\frac{q+1}{2}$\\ \hline \end{tabular} \label{tab2} \end{center} \end{table} \end{Lemma} \section{Some Vector Spaces of Polynomials} Let ${\mathbf P}=\langle(x_0,x_1,x_2)\rangle$ be a point of $PG(2,q)$ and let $\ell=\langle(y_0,y_1,y_2)^{\top}\rangle$ be a line of $PG(2,q)$. We have ${\mathbf P}\in \ell$ if and only if $x_0y_0+x_1y_1+x_2y_2 = 0$. So the $(\ell,{\mathbf P})$-entry of ${\mathbf A}$ is given by \begin{displaymath} ({\mathbf A})_{\ell,{\mathbf P}} = 1 - (x_0y_0+x_1y_1+x_2y_2)^{q-1} =\begin{cases} 1, & \text{if} \; x_0y_0+x_1y_1+x_2y_2 =0,\\ 0, & \text{otherwise}. \end{cases} \end{displaymath} Now we define a $(q^3-1)\times q^3$ matrix $\mathbf{S}$, whose rows are indexed by vectors ${\bf y}=(y_0,y_1,y_2)\in{\mathbb F}_q^3\setminus\{(0,0,0)\}$, whose columns are indexed by vectors ${\bf x}=(x_0,x_1,x_2)\in {\mathbb F}_q^3$, and the $({\bf y},{\bf x})$-entry is equal to $1 - (x_0y_0+x_1y_1+x_2y_2)^{q-1}$. Similar to the partition of ${\mathbf A}$ defined in Section 1, we partition $\mathbf{S}$ into the following form \begin{displaymath} \mathbf{S}=\left( \begin{array}{ccc} \mathbf{S}_{11}& \mathbf{S}_{12} & \mathbf{S}_{13} \\ \mathbf{S}_{21}& \mathbf{S}_{22} & \mathbf{S}_{23} \\ \mathbf{S}_{31}& \mathbf{S}_{32} & \mathbf{S}_{33} \\ \end{array} \right), \end{displaymath} where the rows of $\mathbf{S}_{11}$, $\mathbf{S}_{21}$, and $\mathbf{S}_{31}$ are labeled by vectors $(y_0,y_1,y_2)\in {\mathbb F}_q^3\setminus\{(0,0,0)\}$ such that $y_1^2-4y_0y_2=0$, $y_1^2-4y_0y_2\in\;\not\!\!\Box_q$, and $y_1^2-4y_0y_2\in\;\Box_q$, respectively, and the columns of $\mathbf{S}_{11}$, $\mathbf{S}_{12}$, and $\mathbf{S}_{13}$ are labeled by vectors $(x_0,x_1,x_2)\in {\mathbb F}_q^3$ such that $x_1^2-x_0x_2=0$, $x_1^2-x_0x_2\in\;\not\!\!\Box_q$, and $x_1^2-x_0x_2\in\;\Box_q$, respectively. We also define four more submatrices of $\mathbf{S}$: \begin{displaymath} \mathbf{S}_{nonsec} =\left(\begin{array}{ccc} \mathbf{S}_{11} & \mathbf{S}_{12} & \mathbf{S}_{13} \\ \mathbf{S}_{21} & \mathbf{S}_{22} & \mathbf{S}_{23} \\ \end{array}\right),\; \mathbf{S}_{sec}=\left(\begin{array}{ccc} \mathbf{S}_{31} & \mathbf{S}_{32} & \mathbf{S}_{33}\end{array}\right), \end{displaymath} \begin{displaymath} \mathbf{S}_{sk}=\left(\begin{array}{ccc} \mathbf{S}_{21} & \mathbf{S}_{22} & \mathbf{S}_{23}\end{array}\right),\;\text{and}\; \mathbf{S}_{T}=\left(\begin{array}{ccc}\mathbf{S}_{11}& \mathbf{S}_{12} & \mathbf{S}_{13} \end{array}\right). \end{displaymath} \begin{Lemma}\label{equal} Use the above notation, we have \begin{itemize} \item[(i)] $rank_p{\mathbf A} = rank_p\mathbf{S}$. \item[(ii)] $rank_p\mathbf{S}_{i j} = rank_p\mathbf{A}_{i j}$ for $1\le i, j \le 3$, except $(i,j)=(2,1)$. \item[(iii)] $rank_p{\mathbf A}_{nonsec} = rank_p \mathbf{S}_{nonsec}$, $rank_p {\mathbf A}_{sec} = rank_p \mathbf{S}_{sec}$. \item[(iv)] $rank_p\mathbf{S}_{sk}=rank_p{\mathbf A}_{sk}$. \end{itemize} \end{Lemma} { \noindent{\bf Proof:}\quad } Note that the rows of ${\mathbf S}$ are indexed by the vectors $(y_0,y_1,y_2)\in{\mathbb F}_q^3\setminus\{(0,0,0)\}$ whose transposes represent the lines of $PG(2,q)$. Assume that the first column of $S$ is indexed by $(0,0,0)$. Hence each entry of the first column of ${\mathbf S}$ is $1$. Since each line contains $q+1$ points and each point can be represented by $q-1$ different non-zero vectors, we see that the sum of all the columns of ${\mathbf S}$ is a $q^2$ multiple of an all one column vector of the proper size, which is a zero column vector over ${\mathbb F}_q$. This implies that the first column of ${\mathbf S}$ is a linearly combination of all the other columns of ${\mathbf S}$ over ${\mathbb F}_q$. Hence, the matrix ${\mathbf S}^{'}$ obtained by deleting the first column of ${\mathbf S}$ has the same $p$-rank as ${\mathbf S}$. By deleting duplicate rows and columns of ${\mathbf S}^{'}$ and permuting the rows and columns of the resulting matrix, we can obtain the matrix ${\mathbf A}$. This shows that ${\mathbf S}$, ${\mathbf S}^{'}$, and ${\mathbf A}$ have the same $p$-rank. So $(i)$ follows. It is clear that $rank_p{\mathbf S}_{ij} = rank_p{\mathbf A}_{ij}$ for $(i,j)=(1,3)$, $(2,2)$, $(2,3)$, $(3,2)$, and $(3,3)$. Note that each tangent line contains a point of $\mathcal{O}$ and each point can be represented by $q-1$ non-zero vectors. So the sum of all the columns of $S_{11}$ is a zero column over ${\mathbb F}_q$, which indicates that the first column of ${\mathbf S}_{11}$ is a linear combination of all other columns of ${\mathbf S}_{11}$. Thus, by deleting the first column of ${\mathbf S}_{11}$ and the duplications of rows and columns of ${\mathbf S}_{11}$ and permuting the rows and columns of the resulting matrix, we get ${\mathbf A}_{11}$. Hence, $rank_p{\mathbf S}_{11} = rank_p{\mathbf A}_{11}$. Similarly, the first column of ${\mathbf S}_{31}$ is a linear combination of all the other columns of ${\mathbf S}_{31}$ since, again, the sum of all the columns of ${\mathbf S}_{31}$ is a zero column by noting that each secant line contains $2$ points of $\mathcal{O}$. Thus, by deleting the first column of ${\mathbf S}_{31}$ and the duplications of rows and columns of ${\mathbf S}_{31}$ and permuting the rows and columns of the resulting matrix, we get ${\mathbf A}_{31}$. Hence, $rank_p{\mathbf S}_{31} = rank_p{\mathbf A}_{31}$. It is also clear that $rank_p{\mathbf S}_{21}=1$ and $rank_p{\mathbf A}_{21} =0$. So $(ii)$ is proved. The proofs of $(iii)$ and $(iv)$ are essentially the same as the proofs of $(i)$ and $(ii)$. We omit the detail. \qed\medskip\par \begin{Lemma}\label{in_1} $rank_p \mathbf{S}_{sk} = \binom{p+1}{2}^e-q$. \end{Lemma} { \noindent{\bf Proof:}\quad } Note that the rows of ${\mathbf A}_{nonsec}$ indexed by the tangent lines are linear independent since ${\mathbf A}_{11}$ is a permutation matrix. When deleting the rows of ${\mathbf A}_{nonsec}$ indexed by the tangent lines, the $p$-rank of the matrix decreases by $q+1$. This implies that $rank_p{\mathbf A}_{sk}=\binom{p+1}{2}^e-q$ by $(i)$ of Theorem~\ref{sub_1}. From $(iv)$ of Lemma~\ref{equal}, the lemma follows. \qed\medskip\par Now we define \begin{displaymath} \begin{array}{lll} \mathcal{Z}(Sk) & =&\{(y_0,y_1,y_2)\in {\mathbb F}_q^3\mid y_1^2-4y_0y_2\in \;\not\!\!\Box_q\},\\ \mathcal{Z}(Se) & =&\{(y_0,y_1,y_2)\in {\mathbb F}_q^3\mid y_1^2-4y_0y_2\in \;\Box_q\},\\ \mathcal{Z}(T) &=&\{(y_0,y_1,y_2)\in {\mathbb F}_q^3\mid y_1^2-4y_0y_2 = 0\}\setminus\{(0,0,0)\}.\\ \end{array} \end{displaymath} Also \begin{displaymath} \begin{array}{lll} \mathcal{Z}(I) &=&\{(x_0,x_1,x_2)\in {\mathbb F}_q^3\mid x_1^2-x_0x_2\in\;\not\!\!\Box_q\},\\ \mathcal{Z}(E) &=&\{(x_0,x_1,x_2)\in {\mathbb F}_q^3\mid x_1^2-x_0x_2\in \;\Box_q\},\\ \mathcal{Z}(\mathcal{O})&=&\{(x_0,x_1,x_2)\in {\mathbb F}_q^3\mid x_1^2- x_0x_2 = 0\}.\\ \end{array} \end{displaymath} Let ${\mathbb F}_q[X_0,X_1,X_2]$ be the vector space of polynomials in three indeterminates over ${\mathbb F}_q$ and $F_{q-1}[X_0,X_1,X_2]$ be the subspace of ${\mathbb F}_q[X_0,X_1,X_2]$ over ${\mathbb F}_q$ consisting of the homogeneous polynomials of degree $q-1$, together with $0$. Define the following subspaces of ${\mathbb F}_q[X_0,X_1,X_2]$ over ${\mathbb F}_q$: \begin{displaymath} \begin{array}{llll} \mathcal{M}&=& \left\langle 1 - (y_0X_0+y_1X_1 + y_2X_2)^{q-1}\mid (y_0,y_1,y_2)\in {\mathbb F}_q^3\right\rangle,\\ \mathcal{M}^{Sk} &=& \left\langle1 - (y_0X_0+y_1X_1 + y_2X_2)^{q-1}\mid (y_0,y_1,y_2)\in \mathcal{Z}(Sk)\right\rangle,\\ \mathcal{M}^{Se}&=& \left\langle1 - (y_0X_0+y_1X_1 + y_2X_2)^{q-1}\mid (y_0,y_1,y_2)\in \mathcal{Z}(Se)\right\rangle,\\ \mathcal{M}^{T}&= &\left\langle1 - (y_0X_0+y_1X_1 + y_2X_2)^{q-1}\mid (y_0,y_1,y_2)\in \mathcal{Z}(T)\right\rangle.\\ \end{array} \end{displaymath} Given a vector space $M$ over ${\mathbb F}_q$, we use $dim_q M$ to denote its dimension over ${\mathbb F}_q$. If $\mathbf{B}$ is a matrix over ${\mathbb F}_q$, then we use row$(\mathbf{B})$ to denote the span of the rows of $\mathbf{B}$ over ${\mathbb F}_q$. \begin{Lemma}\label{dim_1} Use the above notation, \begin{itemize} \item[(i)] $dim_q \mathcal{M}^{Sk} = \binom{p+1}{2}^e-q$, \item[(ii)] $dim_q\mathcal{M}^{Se} = \binom{p+1}{2}^e$, \item[(iii)] $dim_q \mathcal{M}^{T} = q+1$. \end{itemize} \end{Lemma} { \noindent{\bf Proof:}\quad } Define $$\psi_0: \mathcal{M}^{Sk}\rightarrow \text{row}({\mathbf{S}_{sk}})$$ by $$g(X_0,X_1,X_2)\mapsto \left(g(x_0,x_1,x_2): (x_0,x_1,x_2)\in {\mathbb F}_q^3\right).$$ Note that $\psi_0([1-(y_0X_0+y_1X_1+y_2X_2)^{q-1}])$ is simply the row of $\mathbf{S}_{sk}$ indexed by $(y_0,y_1,y_2)$ with $y_1^2-4y_0y_2\in \;\not\!\!\Box_q$. Let $$g(X_0,X_1,X_2) =\displaystyle\sum_{(y_0,y_1,y_2)\in \mathcal{Z}(Sk)}a_{y_0,y_1,y_2}\left[1-(y_0X_0+y_1X_1+y_2X_2)^{q-1}\right] \in \mathcal{M}^{Sk}.$$ Then \begin{displaymath} \begin{array}{lll} \psi_0(g(X_0,X_1,X_2)) & = & \displaystyle\sum_{(y_0,y_1,y_2)\in \mathcal{Z}(sk)}a_{y_0,y_1,y_2}\mathbf{S}_{y_0,y_1,y_2}^{sk}\\ {}& = & \displaystyle\sum_{(y_0,y_1,y_2)\in \mathcal{Z}(sk)}a_{y_0,y_1,y_2}\psi_0([1-(y_0X_0+y_1X_1+y_2X_2)^{q-1}])\in\text{row}(\mathbf{S}_{sk}),\\ \end{array} \end{displaymath} where $\mathbf{S}_{y_0,y_1,y_2}^{sk}$ is the row vector of $\mathbf{S}_{sk}$ indexed by $(y_0,y_1,y_2)$. The first equality above comes from the definition of $\psi_0$. So $\psi_0$ is a surjective ${\mathbb F}_q$-linear map. Now assume that $g(X_0,X_1,X_2)\in\text{Ker}(\psi_0)$. Then $g(x_0,x_1,x_2) = 0 $ for all $(x_0,x_1,x_2)\in {\mathbb F}_q^3$. So $g(X_0,X_1,X_2) = 0$. Thus, we proved that $$\mathcal{M}^{Sk}\cong \text{row}(\mathbf{S}_{sk})$$ as vector spaces over ${\mathbb F}_q$. Part $(i)$ of the lemma follows immediately from Lemma~\ref{in_1}. The proofs of part $(ii)$ and $(iii)$ are similar. We need only check that the maps $$\psi_1: \mathcal{M}^{Se}\rightarrow \text{row}({\mathbf{S}_{se}})$$ defined by $$g(X_0,X_1,X_2)\mapsto \left(g(x_0,x_1,x_2): (x_0,x_1,x_2)\in {\mathbb F}_q^3\right)$$ and $$\psi_2: \mathcal{M}^{T}\rightarrow \text{row}({\mathbf{S}_{T}})$$ defined by $$g(X_0,X_1,X_2)\mapsto \left(g(x_0,x_1,x_2): (x_0,x_1,x_2)\in {\mathbb F}_q^3\right)$$ are both ${\mathbb F}_q$-isomorphisms. Since they are the same as the proof of $(i)$, we omit the details. \qed\medskip\par \begin{Theorem}\label{in_2} As ${\mathbb F}_q$-spaces, \begin{itemize} \item[(i)]$\mathcal{M}^{Sk}\cong \text{row}(\mathbf{S}_{22})$, \item[(ii)]$\mathcal{M}^{Se}\cong \text{row}(\mathbf{S}_{33})$, \item[(iii)] $\mathcal{M}^{Sk}\cong \text{row}(\mathbf{S}_{23})$, \item[(iv)] $\mathcal{M}^{T}/\langle\mathbf{J}(X_0,X_1,X_2)\rangle\cong \text{row}(\mathbf{S}_{13})$, where $\langle\mathbf{J}(X_0,X_1,X_2)\rangle$ is the $1$-dimensional subspace generated by the polynomial $$\mathbf{J}(X_0,X_1,X_2) = 1+\displaystyle\sum_{(y_0,y_1,y_2)\in\mathcal{Z}(T)}(y_0X_0+y_1X_1+x_2Y_2)^{q-1}$$ over ${\mathbb F}_q$. \end{itemize} \end{Theorem} { \noindent{\bf Proof:}\quad } Define the map $$\gamma: \mathcal{M}^{sk}\rightarrow \text{row}(\mathbf{S}_{22})$$ by $$g(X_0,X_1,X_2)\mapsto \left(g(x_0,x_1,x_2): (x_0,x_1,x_2)\in \mathcal{Z}(I)\right).$$ We can show that $\gamma$ is a well-defined surjective ${\mathbb F}_q$-linear map by applying arguments similar to the ones of Lemma~\ref{dim_1}. Now let $h(X_0,X_1,X_2)\in\text{Ker}(\gamma)$. Then \begin{equation} \begin{array}{lll} h(x_0,x_1,x_2) &= & \displaystyle\sum_{(y_0,y_1,y_2)\in \mathcal{Z}(Sk)}a_{y_0,y_1,y_2}\left[1-(x_0y_0+x_1y_1+x_2y_2)^{q-1}\right]\\ {} & = & \displaystyle\sum_{(y_0,y_1,y_2)\in\mathcal{Z}(Sk)}a_{y_0,y_1,y_2} - \displaystyle\sum_{(y_0,y_1,y_2)\in\mathcal{Z}(Sk)}a_{y_0,y_1,y_2}(x_0y_0+x_1y_1+x_2y_2)^{q-1}\\ {}& = & 0 \end{array} \end{equation} for all $(x_0,x_1,x_2)\in\mathcal{Z}(I)$, where $a_{y_0,y_1,y_2}\in {\mathbb F}_q$. Note that $(0,0,0)\notin\mathcal{Z}(I)$. Let ${\mathbf P}=\langle(x_0,x_1,x_2)\rangle$ be any internal point and Sk$_{{\mathbf P}}$ be the set of $\frac{q+1}{2}$ skew lines through ${\mathbf P}$. We denote the set of row vectors whose transposes represent the skew lines through ${\mathbf P}$ by $\mathcal{Z}(Sk_{{\mathbf P}})$. Then $(1)$ gives \begin{displaymath} h(x_0,x_1,x_2) = \displaystyle\sum_{(y_0,y_1,y_2)\in\mathcal{Z}(Sk_{{\mathbf P}})} a_{y_0,y_1,y_2} = 0. \end{displaymath} Hence \begin{equation} \displaystyle\sum_{{\mathbf P}=\langle(x_0,x_1,x_2)\rangle\in I}h(x_0,x_1,x_2) = \displaystyle\sum_{{\mathbf P}=\langle(x_0,x_1,x_2)\rangle\in I}\displaystyle\sum_{(y_0,y_1,y_2)\in\mathcal{Z}(Sk_{{\mathbf P}})}a_{y_0,y_1,y_2} = 0. \end{equation} Since both $\langle(y_0,y_1,y_2)^{\top}\rangle$ and $\langle(x_0,x_1,x_2)\rangle$ are represented by $(q-1)$ different non-zero vectors, we see that $|\mathcal{Z}(Sk_{{\mathbf P}})|=\frac{(q+1)(q-1)}{2}$ by the last row of Table~\ref{tab2} and $|\mathcal{Z}(I)|=\frac{q(q-1)^2}{2}$ by the Table~\ref{table:points}. So $(2)$ can be written as \begin{displaymath} \begin{array}{lll} \displaystyle\sum_{{\mathbf P}=\langle(x_0,x_1,x_2)\rangle\in I}\displaystyle\sum_{(y_0,y_1,y_2)\in\mathcal{Z}(Sk_{{\mathbf P}})}a_{y_0,y_1,y_2} &=& \frac{q+1}{2}\cdot\left(\displaystyle\sum_{(y_0,y_1,y_2)\in\mathcal{Z}(Sk)}a_{y_0,y_1,y_2}\right) \\ {} &=& 0. \end{array} \end{displaymath} As $p\nmid\frac{q+1}{2}$ for odd $p$, we must have \begin{displaymath} \begin{array}{lll} \displaystyle\sum_{(y_0,y_1,y_2)\in\mathcal{Z}(Sk)}a_{y_0,y_1,y_2}& = & 0. \end{array} \end{displaymath} Thus $h(X_0,X_1,X_2)\in F_{q-1}[X_0,X_1,X_2]$ and $$h(x_0,x_1,x_2) = \displaystyle\sum_{(y_0,y_1,y_2)\in\mathcal{Z}(Sk)}a_{y_0,y_1,y_2}(x_0y_0+x_1y_1+x_2y_2)^{q-1} = 0 $$ for each $(x_0,x_1,x_2)\in\mathcal{Z}(I)$. Hence $h(X_0,X_1,X_2) = 0$ by $(ii)$ of Lemma~\ref{new}. We have proved that $\gamma$ is an ${\mathbb F}_q$-isomorphism. So $(i)$ follows. Part $(ii)$ can be proved in the same fashion. Consider the map $$\beta: \mathcal{M}^{Se}\rightarrow \text{row}(\mathbf{S}_{33})$$ by $$g(X_0,X_1,X_2)\mapsto \left(g(x_0,x_1,x_2): (x_0,x_1,x_2)\in \mathcal{Z}(E)\right).$$ It is easy to see that $\beta$ is a well-defined surjective ${\mathbb F}_q$-linear map. Let $h(X_0,X_1,X_2)\in\text{Ker}(\beta)$. Then \begin{equation} \begin{array}{lll} h(x_0,x_1,x_2) &= & \displaystyle\sum_{(y_0,y_1,y_2)\in \mathcal{Z}(Se)}a_{y_0,y_1,y_2}\left[1-(x_0y_0+x_1y_1+x_2y_2)^{q-1}\right]\\ {} & = & \displaystyle\sum_{(y_0,y_1,y_2)\in\mathcal{Z}(Se)}a_{y_0,y_1,y_2} - \displaystyle\sum_{(y_0,y_1,y_2)\in\mathcal{Z}(Se)}a_{y_0,y_1,y_2}(x_0y_0+x_1y_1+x_2y_2)^{q-1}\\ {}& = & 0 \end{array} \end{equation} for all $(x_0,x_1,x_2)\in\mathcal{Z}(E)$, where $a_{y_0,y_1,y_2}\in {\mathbb F}_q$. Note that $(0,0,0)\notin\mathcal{Z}(E)$. Let ${\mathbf P}=\langle(x_0,x_1,x_2)\rangle$ be any external point and Se$_{{\mathbf P}}$ be the set of secant lines through ${\mathbf P}$. We denote the set of different vectors $(y_0,y_1,y_2)$ whose transposes represent the secant lines through ${\mathbf P}$ by $\mathcal{Z}(Se_{{\mathbf P}})$. Then $(3)$ gives \begin{displaymath} h(x_0,x_1,x_2) = \displaystyle\sum_{(y_0,y_1,y_2)\in\mathcal{Z}(Se_{{\mathbf P}})} a_{y_0,y_1,y_2} = 0. \end{displaymath} Hence \begin{equation}\label{sum_2} \displaystyle\sum_{{\mathbf P}=\langle(x_0,x_1,x_2)\rangle\in E}h(x_0,x_1,x_2) = \displaystyle\sum_{{\mathbf P}=\langle(x_0,x_1,x_2)\rangle\in E}\displaystyle\sum_{(y_0,y_1,y_2)\in\mathcal{Z}(Se_{{\mathbf P}})}a_{y_0,y_1,y_2} = 0. \end{equation} Since $|\mathcal{Z}(Se_{{\mathbf P}})|=\frac{(q-1)^2}{2}$ by the last row of Table~\ref{tab2} and $|\mathcal{Z}(E)|=\frac{q(q+1)(q-1)}{2}$ by the Table~\ref{table:points}, we see that $(\ref{sum_2})$ can be written as \begin{displaymath} \begin{array}{llll} \displaystyle\sum_{{\mathbf P}=\langle(x_0,x_1,x_2)\rangle\in E}\displaystyle\sum_{(y_0,y_1,y_2)\in\mathcal{Z}(Se_{{\mathbf P}})}a_{y_0,y_1,y_2} &= &\frac{q-1}{2}\cdot\left(\displaystyle\sum_{(y_0,y_1,y_2)\in\mathcal{Z}(Se)}a_{y_0,y_1,y_2}\right)\\ {} & = & 0.\\ \end{array} \end{displaymath} As $p\nmid\frac{q-1}{2}$ for odd $p$, we must have \begin{displaymath} \begin{array}{lll} \displaystyle\sum_{(y_0,y_1,y_2)\in\mathcal{Z}(Se)}a_{y_0,y_1,y_2} & = & 0. \end{array} \end{displaymath} Thus $h(X_0,X_1,X_2) \in F_{q-1}[X_0,X_1,X_2]$ and \begin{displaymath} \begin{array}{lll} h(x_0,x_1,x_2)& = &\displaystyle\sum_{(y_0,y_1,y_2)\in\mathcal{Z}(Se)}a_{y_0,y_1,y_2}(x_0y_0+x_1y_1+x_2y_2)^{q-1} \\ {} & = & 0. \end{array} \end{displaymath} for all $(x_0,x_1,x_2)\in\mathcal{Z}(E)$. Hence $h(X_0,X_1,X_2) = 0$ by $(i)$ of Lemma~\ref{new}. So $(ii)$ is proved. The proof of $(iii)$ is essentially the same as the proof of $(i)$, so we omit the details. To prove $(iv)$, consider the map $$\eta: \mathcal{M}^{T}\rightarrow \text{row}(\mathbf{S}_{13})$$ by $$g(X_0,X_1,X_2)\mapsto \left(g(x_0,x_1,x_2): (x_0,x_1,x_2)\in \mathcal{Z}(E)\right).$$ Again, it is easy to see that $\eta$ is a well-defined surjective ${\mathbb F}_q$-linear map. Let $h(X_0,X_1,X_2)\in\text{Ker}(\eta)$. Then \begin{equation} \begin{array}{lll} h(x_0,x_1,x_2) &= & \displaystyle\sum_{(y_0,y_1,y_2)\in \mathcal{Z}(T)}a_{y_0,y_1,y_2}\left[1-(x_0y_0+x_1y_1+x_2y_2)^{q-1}\right]\\ {} & = & \displaystyle\sum_{(y_0,y_1,y_2)\in\mathcal{Z}(T)}a_{y_0,y_1,y_2} - \displaystyle\sum_{(y_0,y_1,y_2)\in\mathcal{Z}(T)}a_{y_0,y_1,y_2}(x_0y_0+x_1y_1+x_2y_2)^{q-1}\\ {}& = & 0 \end{array} \end{equation} for all $(x_0,x_1,x_2)\in\mathcal{Z}(E)$, where $a_{y_0,y_1,y_2}\in {\mathbb F}_q$. Note that $(0,0,0)\notin\mathcal{Z}(E)$. Let ${\mathbf P}=\langle(x_0,x_1,x_2)\rangle$ be any external point. We use $\mathcal{Z}(T_{{\mathbf P}})$ to denote the row vectors whose transposes represent either of two tangent lines through ${\mathbf P}$. Then $(5)$ simplifies to \begin{displaymath} \begin{array}{lll} h(x_0,x_1,x_2) &=& \displaystyle\sum_{(y_0,y_1,y_2)\in \mathcal{Z}(T_{{\mathbf P}})}a_{y_0,y_1,y_2}\\ {}& =& 0.\\ \end{array} \end{displaymath} Let $E_{\ell}$ denote the set of external points on a given tangent line $\ell$ and $\langle\ell\rangle$ be the $1$-dimensional subspace over ${\mathbb F}_q$ generated by the row vector whose transpose represents $\ell$. Since $T_{{\mathbf P}}$ consists of the two tangent lines through ${\mathbf P}$ (the second row of Table~\ref{tab2}), then, in the multiset $\bigcup_{{\mathbf P}\in E_{\ell}}T_{{\mathbf P}}$, each tangent line other than $\ell$ appears exactly once, and $\ell$ appears exactly $q$ times. Hence, \begin{displaymath} \begin{array}{llll} \displaystyle\sum_{{\mathbf P}=\langle(x_0,x_1,x_2)\rangle\in E_{\ell}}h(x_0,x_1,x_2) &= &\displaystyle\sum_{{\mathbf P}=\langle(x_0,x_1,x_2)\rangle\in E_{\ell}}\displaystyle\sum_{(y_0,y_1,y_2)\in \mathcal{Z}(T_{{\mathbf P}})}a_{y_0,y_1,y_2} \\ {} & = & q\cdot\left(\displaystyle\sum_{(y_0,y_1,y_2)\in\langle\ell\rangle}a_{y_0,y_1,y_2}\right) + \displaystyle\sum_{(y_0,y_1,y_2)\in\mathcal{Z}(T)\setminus\langle\ell\rangle}a_{y_0,y_1,y_2}\\ {} & = & \displaystyle\sum_{(y_0,y_1,y_2)\in\mathcal{Z}(T)\setminus\langle\ell\rangle}a_{y_0,y_1,y_2} \\ {} & = & 0. \end{array} \end{displaymath} Hence for any two different tangent lines $\ell_1$ and $\ell_2$, we have \begin{displaymath} \begin{array}{lll} \displaystyle\sum_{(y_0,y_1,y_2)\in\mathcal{Z}(T)\setminus\langle\ell_1\rangle}a_{y_0,y_1,y_2} &=& \displaystyle\sum_{(y_0,y_1,y_2)\in\mathcal{Z}(T)\setminus\langle\ell_2\rangle}a_{y_0,y_1,y_2}, \end{array} \end{displaymath} which indicates that \begin{displaymath} \begin{array}{lll} \displaystyle\sum_{(y_0,y_1,y_2)\in \langle\ell_1\rangle} a_{y_0,y_1,y_2} &= & \displaystyle\sum_{(y_0,y_1,y_2)\in \langle\ell_2\rangle} a_{y_0,y_1,y_2} = L\in {\mathbb F}_q\\ \end{array} \end{displaymath} for any two tangent lines $\ell_1$ and $\ell_2$. Thus, \begin{equation}\label{eq6} \begin{array}{llllll} h(X_0,X_1,X_2) & = & \displaystyle\sum_{(y_0,y_1,y_2)\in\mathcal{Z}(T)}a_{y_0,y_1,y_2} - \displaystyle\sum_{(y_0,y_1,y_2)\in\mathcal{Z}(T)}a_{y_0,y_1,y_2}(X_0y_0+X_1y_1+X_2y_2)^{q-1}\\ {}& = & \displaystyle\sum_{\ell\in T}\displaystyle\sum_{(y_0,y_1,y_2)\in\langle\ell\rangle}a_{y_0,y_1,y_2} - \displaystyle\sum_{\ell\in T}\left(\displaystyle\sum_{(y_0,y_1,y_2)\in \langle\ell\rangle}a_{y_0,y_1,y_2}\right)(y_0X_0+y_1X_1+y_2X_2)^{q-1}\\ {} & = & L\cdot(q+1) - L\cdot\left[\displaystyle\sum_{\langle(y_0,y_1,y_2)^T\rangle\in T}(X_0y_0+X_1y_1+X_2y_2)^{q-1}\right]\\ {}&=& L + L\cdot\left[(q-1)\cdot\displaystyle\sum_{\langle(y_0,y_1,y_2)^T\rangle\in T}(X_0y_0+X_1y_1+X_2y_2)^{q-1}\right]\\ {} & = & L\cdot\left[ 1 + \displaystyle\sum_{(y_0,y_1,y_2)\in \mathcal{Z}(T)}(X_0y_0+X_1y_1+X_2y_2)^{q-1}\right].\\ \end{array} \end{equation} The second and the last equalites in (\ref{eq6}) hold since for any two vectors $(y_0,y_1,y_2),\;(y_0^{'},y_1^{'},y_2^{'})\in\langle\ell\rangle$, $(X_0y_0+X_1y_1+X_2y_2)^{q-1}=(X_0y_0^{'}+X_1y_1^{'}+X_2y_2^{'})^{q-1}$. So $(iv)$ follows immediately. \qed\medskip\par \begin{Lemma}\label{you1} Let $\mathbf{S}_{ij}$'s with $1\le i, j \le 3$ be the submatrices of $\mathbf{S}$. Then \begin{itemize} \item[(i)] $rank_p(\mathbf{S}_{11}) = q+1$, \item[(ii)] $rank_p(\mathbf{S}_{12}) = 0$, \item[(iii)] $rank_p(\mathbf{S}_{13}) = rank_p(\mathbf{S}_{31}) = q$, \item[(iv)] $rank_p(\mathbf{S}_{22}) = rank_p(\mathbf{S}_{23}) = rank_p(\mathbf{S}_{32}) = \binom{p+1}{2}^e - q$, \item[(v)] $rank_p(\mathbf{S}_{33}) = \binom{p+1}{2}^e$. \end{itemize} \end{Lemma} { \noindent{\bf Proof:}\quad } $(i)$ and $(ii)$ are clear. By $(iv)$ of Theorem~\ref{in_2}, we have $rank_p(\mathbf{S}_{13})=q$. Since $\mathbf{S}_{31}$ can be obtained by permuting the rows and columns of $\mathbf{S}_{13}$, then $rank_p(\mathbf{S}_{31}) = q$. $(iv)$ follows from $(i)$ and $(iii)$ of Theorem~\ref{in_2} and $(ii)$ of Lemma~\ref{dim_1}. $(iv)$ follows from $(ii)$ of Theorem~\ref{in_2} and the fact that ${\mathbf S}_{32}$ can be obtained by permuting the rows and columns of ${\mathbf S}_{23}$. Finally, $(v)$ follows from $(ii)$ of Lemma~\ref{dim_1} and $(ii)$ of Theorem~\ref{in_2}. \qed\medskip\par \noindent{{\bf Proof of Theorem~\ref{you}:}} The theorem follows immediately from Corollary~\ref{equal} and Lemma~\ref{you1}. \qed\medskip\par
\section{Introduction} With this paper, we launch a project to classify integrable classes of surfaces. These are classes of surfaces whose Gauss--Mainardi--Codazzi equations are integrable in the sense of soliton theory. Our long-term goals include obtaining lists of integrable classes as complete as computing resources permit, clarifying their mutual relations, and identifying known subcases. Our immediate goal is to demonstrate that the task is feasible and worth doing. The classical geometry of immersed surfaces in the Euclidean space is well known to be closely connected with the modern theory of integrable systems~\cite{R-S}. The Gauss--Weingarten equations of a moving frame $\Psi$ always take the form $$ \numbered\label{lin problem} \Psi_x = A \Psi, \quad \Psi_y = B \Psi. $$ where $A,B$ are appropriate matrix functions. Integrability conditions of~\eqref{lin problem} are called the Gauss--Mainardi--Codazzi equations and take the form of a {\it zero curvature representation} $$ \numbered\label{ZCR} A_{y} - B_{x} + [A,B] = 0. $$ Equation~\eqref{ZCR} is invariant under a huge group of {\it gauge transformations} $$ \numbered\label{gauge} A' = S_x S^{-1} + S A S^{-1}, \qquad B' = S_y S^{-1} + S B S^{-1}, $$ induced by linear transformations $\Psi' = S \Psi$ of the frame. Here $S$ is an invertible functional matrix, which can be restricted to take values in the Lie group $G$ associated with the Lie algebra $\mathfrak g$ matrices $A,B$ belong to -- typically $\mathfrak{so}(3)$. The zero curvature representation~\eqref{ZCR} is the key ingredient in the soliton theory~\cite{F-T}, where matrices $A,B$ are additionally assumed to depend on what is called the {\it spectral parameter}. The essential requirement for solitonic integrability is that the spectral parameter cannot be removed by means of the gauge transformation~\eqref{gauge}. Consequently, if the matrices $A,B$ can be modified so that they depend on a nonremovable parameter and still satisfy~\eqref{ZCR}, then the corresponding Gauss--Mainardi--Codazzi equations are considered to be integrable in the sense of soliton theory, and their solutions are known as {\it integrable} or {\it soliton surfaces}~\cite{Sym}. Solitonic integrability can appear only when surfaces are subject to a constraint (such as being pseudospherical etc.). For numerous classical and recent examples see, e.g., the references~\cite{Bob,R-S,S-K} (or~\cite{Fer} in the projective setting). Workable tools to classify such constraints include all the general integrability criteria~\cite{M-S}, which are, however, not immediately applicable to non-evolutionary systems~\cite{M-N-W}. Other methods take advantage of the already known non-parametric zero curvature representation~\eqref{ZCR}, e.g., the method of extended symmetries by Cie\'sli\'nski et al.~\cite{C-ls,C-nls,C-G-S}. In this paper we employ a recent method due to one of us~\cite{spp}. Its essence can be summarized as follows: We attempt to extend the given non-parametric zero curvature representation (a seed) to a power series in terms of the spectral parameter. In the work~\cite{spp}, the relevant computable cohomological obstructions are identified. Two obstacles make this procedure not entirely algorithmic: The parameter-dependent zero curvature representation can exist only in an extension of the Lie algebra $\mathfrak g$ and its jet order (the order of derivatives) can exceed that of the seed. If no obstructions are found, various ways exist to incorporate the true nonremovable parameter. \section{Weingarten surfaces} To be of genuine interest in geometry, the determining constraint on integrable surfaces must be invariant with respect to coordinate changes. The general non-differential invariant constraint is a functional relation $f(p,q) = 0$ between the principal curvatures $p,q$. Such a functional relation is characteristic of Weingarten surfaces, which have been a topic of continuous interest, especially in global differential geometry~\cite{Hop,Vos,S-K,Lop} and computer graphics~\cite{vBrunt}. Well known to be integrable is the class of {\it linear Weingarten surfaces}~\cite{Dar,R-S}, characterized by a linear relation $$ \numbered\label{lws} a k + b h + c = 0, \qquad a,b,c = `const $$ between the Gauss curvature $k = p q$ and the mean curvature $h = \frac 12 (p + q)$ (not to be mixed with a linear relation between the principal curvatures~\cite{K-S,Lop}). Other integrable classes of Weingarten surfaces that sporadically occur in the literature all have a differential defining relation (e.g., the Hazzidakis equation of the Bonnet surfaces~\cite{Bob,B-E,Bon}; a harmonicity condition of Schief's~\cite{Sch} generalized linear Weingarten surfaces) or the class is not determined by the functional relation $f(p,q) = 0$ alone (e.g.,~\cite{C-F-G}). So far, nothing contradicts the conjecture of Finkel~\cite[Conjecture~3.4]{Fin} and Wu~\cite{Wu} that the only functional relation $f(p,q) = 0$ to determine an integrable class of Weingarten surfaces is the linear relation~\eqref{lws}. Supporting arguments include Wu's~\cite{Wu} proof of non-existence of an $\mathfrak{so}(3)$-valued zero-curvature representation depending only on $x$-derivatives. Finkel's~\cite{Fin} argument roots in an unsuccessful search for higher-order symmetries and a (disputable, see~\cite[\S2]{M-N-W}) conjecture that integrability implies the existence of a local higher-order symmetry (actually the infinite hierarchy can be nonlocal, see also~\cite[\S1.4.4.2]{M-S}). Nevertheless, the main result of the present paper asserts that the simple relation $$ \numbered\label{nws1} \frac 1p - \frac 1q = `const $$ between the main curvatures $p,q$, determines an integrable class of Weingarten surfaces. The associated nonlinear partial differential equation~\eqref{z} has a parameter-dependent zero curvature representation~\eqref{z:zcr} (outside the class considered in~\cite{Wu}), a third-order symmetry~\eqref{z:sym} (missed in~\cite{Fin}), and a recursion operator~\eqref{z:ro}. Paradoxically enough, surfaces satisfying relation~\eqref{nws1} were not unknown to nine\-teenth century geometers. In view of their knowledge, our integrability result is not an entirely unexpected one. In fact, Ribaucour~\cite{Ri} established that the corresponding focal surfaces (evolutes) have a constant Gaussian curvature $k < 0$ (are pseudospherical). Conversely, surfaces satisfying equation~\eqref{nws1} are involutes of pseudospherical surfaces. Moreover, the classical Bianchi transformation~\cite{Bia} is nothing but the induced correspondence between the two focal pseudospherical surfaces. Ribaucour's theorems are covered in Darboux~\cite{Dar} and early twentieth-century monographs, such as~\cite{BiaI,Eis,For,Wea}. Later they became obsolete and forgotten as the induced Bianchi relation between pseudo\-spherical surfaces became superseded by the classical B\"acklund transformation (the history is nicely reviewed by Prus and Sym in~\cite[Sect.~4]{P-S}). The first examples of surfaces satisfying relation~\eqref{nws1} also date to the nineteenth century. Lipschitz~\cite{Lip} derived a four-parametric family in terms of elliptic integrals. A particular subcase, the rotation surface of von Lilienthal~\cite{Lil}, is the involute surface of the pseudosphere. The left-hand side of Equation~\eqref{nws1} is equal to the difference of the principal radii of curvature at a point. This geometric quantity has a definite physical meaning, being associated with the {\it interval of Sturm}~\cite{Sturm}, also known as the {\it astigmatic interval\/} or the {\it amplitude of astigmatism\/} or simply the {\it astigmatism}~\cite{Gr-I}. A mirror or a refracting surface satisfying relation~\eqref{nws1} will feature a constant amplitude of astigmatism in the normal directions. In the sequel, surfaces satisfying condition~\eqref{nws1} will be called {\it surfaces of constant astigmatism.} Accordingly, the equation~\eqref{z} to determine the surfaces of constant astigmatism will be called the {\it constant astigmatism equation}. \section{Preliminaries} \label{sect:prelim} We shall consider surfaces $\mathbf r(x,y)$ parametrized by curvature lines. As is well known, the fundamental forms can be written as $$ `I = u^2\,`d x^2 + v^2\,`d y^2, \\ `II = u^2 p\,`d x^2 + v^2 q\,`d y^2, $$ where $p,q$ are the principal curvatures. Coordinates $x,y$ are unique up to arbitrary changes $x = X(x)$, $y = Y(y)$. Let $\Psi = (\mathbf e_1,\mathbf e_2,\mathbf n)$ denote the orthonormal frame, given by $\mathbf e_1 = \mathbf r_x/u$, $\mathbf e_2 = \mathbf r_y/v$, $\mathbf n = \mathbf e_1 \times \mathbf e_2$. The Gauss--Weingarten equations $$ \numbered \label{G-W:so3} \Psi_x = \left(\begin{array}{ccc} \hphantom{-} 0 & -\frac{u_y}{v} & \hphantom{-} u p \\ \hphantom{-} \frac{u_y}{v} & \hphantom{-} 0 & \hphantom{-} 0 \\ -u p & \hphantom{-} 0 & \hphantom{-} 0 \end{array}\right) \Psi, \quad \Psi_y = \left(\begin{array}{ccc} \hphantom{-} 0 & \hphantom{-} \frac{v_x}{u} & \hphantom{-} 0 \\ -\frac{v_x}{u} & \hphantom{-} 0 & \hphantom{-} v q \\ \hphantom{-} 0 & -v q & \hphantom{-} 0 \end{array}\right) \Psi. $$ are easily established. Their integrability conditions are the Gauss equation \begin{equation} \label{GMC-G} u u_{yy} + v v_{xx} - \frac{v}{u} u_x v_x - \frac{u}{v} u_y v_y + u^2 v^2 p q = 0, \end{equation} and the Mainardi--Codazzi equations \begin{equation} (p - q) u_y + u p_y = 0, \quad (q - p) v_x + v q_x = 0. \label{GMC-MC} \end{equation} Consequently, the two~$\mathfrak{so}(3)$ matrices occurring in formulas~\eqref{G-W:so3} constitute a nonparametric zero curvature representation of the Gauss--Mainardi--Codazzi system~\eqref{GMC-G},~\eqref{GMC-MC}. Because of the isomorphism $\mathfrak{so}(3,\mathbb C) \cong \mathfrak{sl}(2,\mathbb C)$, the same zero curvature representation can be alternatively written in terms of $2 \times 2$ matrices $$ \numbered \label{0} A_0 = \left(\begin{array}{cc} \frac{`i u_y}{2v} & -\frac{1}{2} u p \\ \frac{1}{2} u p & -\frac{`i u_y}{2v} \end{array}\right), \qquad B_0 = \left(\begin{array}{cc} -\frac{`i v_x}{2u} & -\frac{1}{2} `i q v \\ -\frac{1}{2} `i q v & \hphantom{-} \frac{`i v_x}{2u} \end{array}\right). $$ Let us impose a constraint~$f(p,q) = 0$. If nontrivial, it can be resolved with respect to one of the curvatures, say $$ \numbered\label{q=F(p)} q = F(p), $$ which we assume henceforth. Then the Gauss--Mainardi--Codazzi system reduces substantially~\cite{vBrunt,Fin,Wu}. In particular, the Mainardi--Codazzi equations~\eqref{GMC-MC} have a general solution $$ u = \frac{u_0}E, \quad v = -v_0 E', \quad q = p - \frac{E}{E'}, $$ where $E = E(p)$ is an arbitrary nonconstant function, $E' = dE/dp$, and and $u_0,v_0$ are functions of $x$ and $y$, respectively, removable by transformation $\tilde x = \int u_0\,dx$, $\tilde y = \int v_0\,dy$. Therefore, we can put $u_0 = -v_0 = 1$ without loss of generality, i.e., $$ \numbered\label{E2uvq} u = \frac1E, \quad v = E', \quad q = p - \frac{E}{E'}. $$ The Gauss equation~\eqref{GMC-G} then becomes \begin{equation} \label{G} p_{yy} = E^3 E'' p_{xx} + 2 \frac{E'}E p_y^2 + E^2 (E E'')' p_x^2 + E E' p^2 - E^2 p. \end{equation} Summarizing, the Gauss--Mainardi--Codazzi system of Weingarten surfaces reduces to the single equation~\eqref{G}. The classification problem considered in this paper is ``for which choices of the function $E(p)$ is the equation~\eqref{G} integrable?'' By substituting~\eqref{E2uvq} into \eqref{0}, we easily obtain a nonparametric zero curvature representation of equation~\eqref{G}, $$ \numbered\label{AB0} A_0 = (\begin{array}{cc} \frac{`i}2 \frac{p_y}{E^2} & -\frac12 \frac p E \\ \frac12 \frac p E & -\frac{`i}2 \frac{p_y}{E^2} \end{array}), \quad B_0 = (\begin{array}{cc} \frac{`i}2 E E'' p_x & \frac{`i}2 (E' p - E) \\ \frac{`i}2 (E' p - E) & -\frac{`i}2 E E'' p_x \end{array}), $$ which will be the starting point of the calculations to follow. \section{Cohomological criteria} Readers not interested in details of the classification method can skip this section and continue to investigation of surfaces of constant astigmatism in Section~\ref{z:sect}. We use the formal theory of partial differential equations, which treats coordinates, unknown functions, and their derivatives as independent quantities. Equations can be conveniently represented as submanifolds in appropriate jet spaces~\cite{B-V-V}. All our considerations being local, we let $J^\infty = J^\infty(\mathbb R^2, \mathbb R)$ denote the space of $\infty$-jets of smooth functions $\mathbb R^2 \to \mathbb R$. The base $\mathbb R^2$ being equipped with coordinates $x,y$, the natural coordinates along fibres of $J^\infty \to \mathbb R^2$ correspond to $p$ and its derivatives. These will be denoted $p_I$, where $I$ stands for a symmetric multiindex in $x,y$ (including the ``empty'' multiindex $\emptyset$ such that $p_\emptyset = p$). The usual total derivatives $$ D_x = \frac\partial{\partial x} + \sum_{I} p_{xI} \frac{\partial}{\partial p_I}, \qquad D_y = \frac\partial{\partial y} + \sum_{I} p_{yI} \frac{\partial}{\partial p_I} $$ can be viewed as acting on smooth functions defined on $J^\infty$ (by definition, a smooth function locally depends on a finite number of coordinates). In $J^\infty$, we consider a submanifold $\mathcal G$ determined by equation~\eqref{G} and all its differential consequences obtained by taking successive total derivatives of both sides of~\eqref{G}. On $\mathcal G$, all derivatives of the form $p_{Jyy}$ become expressible in terms of the others. Therefore, derivatives $p_I$ with $y$ occurring no more than twice in $I$ serve as natural coordinates along the fibres of $\mathcal G \to \mathbb R^2$. Being tangent to~\eqref{G}, the total derivatives admit a restriction to $\mathcal G$. We retain the same notation $D_x,D_y$ for the restricted total derivatives. The essence of the adopted point of view can be summarized as follows: A function $f$ on $J^\infty$ satisfies $f|_{\mathcal G} = 0$ if and only if $f$ is zero as a consequence of equation~\eqref{G}. From now on we assume that all objects (like the matrices $A,B$) are defined on $\mathcal G$. When writing $$ \numbered\label{ZCR'} \left.(D_{y} A - D_{x} B + [A,B])\right|_{\mathcal G} = 0 $$ we mean that the zero curvature condition~\eqref{ZCR} holds as a consequence of equation~\eqref{G}. In what follows, characteristic elements~\cite{M1,M2,Sak} play a crucial role. These are nonabelian analogues of characteristics of conservation laws~\cite{B-V-V}. For instance, the characteristic element of the initial zero curvature representation~\eqref{AB0} is the $\mathfrak{sl}(2,\mathbb C)$-matrix $$ C_0 = (\begin{array}{cc} \frac{`i}2 \frac 1{E^2} & 0 \\ 0 & -\frac{`i}2 \frac 1{E^2} \end{array}). $$ This immediately follows from the fact that $$ D_{y} A_0 - D_{x} B_0 + [A_0,B_0] = C_0 F, $$ where $$ F = p_{yy} - E^3 E'' p_{xx} - 2 \frac{E'}E p_y^2 - E^2 (E E'')' p_x^2 - p^2 E E' - p E^2, $$ so that the Gauss equation~\eqref{G} can be written as $F = 0$. Let $A = A(\lambda)$, $B = B(\lambda)$ be the parametric zero curvature representation sought, $C = C(\lambda)$ the corresponding characteristic element. Besides~\eqref{ZCR'}, they will also satisfy the formula~\cite{M1} $$ \numbered\label{ds} \left. \sum_{I} (-\hat D)_I (\frac{\partial F}{\partial u^k_I} C) \right|_{\mathcal G} = 0, $$ with $I$ running over all symmetric multiindices, including the empty one. Here $\hat D_x = D_x - [A,\adot]$, $\hat D_y = D_y - [B,\adot]$, the other values being obtained by composition, which can be taken in any order since~\eqref{ZCR'} implies that $\hat D_x, \hat D_y$ commute. Characteristic elements of gauge equivalent zero curvature representations are conjugate (similar). This allows us to transform characteristic elements into the normal form with respect to conjugation, namely, the Jordan normal form. Since the matrix $C_0$ above is diagonal, it follows that for $\lambda$ sufficiently close to zero the characteristic element $C(\lambda)$ will be also diagonalizable. However, diagonal matrices have a nontrivial stabilizer $\mathcal S \subset `SL(2,\mathbb C)$ with respect to conjugation, which consists of diagonal matrices $$ (\begin{array}{cc} s & 0 \\ 0 & 1/s \end{array}). $$ Gauge transformations from the group $\mathcal S$ (henceforth $\mathcal S$-transformations) preserve the characteristic elements $C(\lambda)$. Their gauge action on a general $\mathfrak{sl}(2)$-valued zero curvature representation $A,B$ is sufficiently simple: $$ (\begin{array}{cc} a_{11} & a_{12} \\ a_{21} & -a_{11} \end{array}) \mapsto (\begin{array}{cc} \frac{s_x}s + a_{11} & s^2 a_{12} \\ \frac{a_{21}}{s^2} & -\frac{s_x}s - a_{11} \end{array}) $$ and similarly for $B$. Using $\mathcal S$-transformations, one can achieve a unique normal form of matrices $A,B$ as follows: If $a_{12} \ne 0$, then by setting $s = (a_{21}/a_{12})^{1/4}$ we turn $A$ into a symmetric matrix, while in the remaining case $a_{12} = 0$ the zero curvature representation degenerates to a pair of conservation laws~\cite{M2}. In other words, being symmetric is a normal form of nondegenerate zero curvature representations with respect to $\mathcal S$-transformations. Turning back to our original problem, we see that $B_0$ is symmetric, and therefore the nearby matrices $B(\lambda)$ can also be symmetrized by an $\mathcal S$-transformation. A simple calculation shows that, by assuming diagonality of $C(\lambda)$ and symmetricity of $B(\lambda)$, we make the system~\eqref{ds} determined, hence solvable (actually, we fix the gauge). Summarizing, the computation of zero curvature representation has been reduced to solution of the determined system~\eqref{ZCR'}, \eqref{ds} under a suitable choice of normal forms for $C$ and $B$. However, this nonlinear system is still quite difficult to solve even with the help of computer algebra. To linearize the system, the work~\cite{spp} considers Taylor expansions $$ \numbered\label{Tayl} A(\lambda) = \sum_{k = 0} A_k \lambda^k, \quad B(\lambda) = \sum_{k = 0} B_k \lambda^k, \quad C(\lambda) = \sum_{k = 0} C_k \lambda^k, $$ with $A_0,B_0,C_0$ coming from the initial parameterless zero curvature representation~\eqref{0}. The condition of zero curvature for $A(\lambda),B(\lambda)$ implies an infinite sequence of conditions of zero curvature for block triangular matrices $$ \numbered\label{barbar AB} A^{[m]} = (\begin{array}{cccc} A_0 & 0 & \dots & 0 \\ A_1 & A_0 & \ddots & \vdots \\ \vdots & \ddots & \ddots & 0 \\ A_m & \dots & A_1 & A_0 \end{array}), \quad B^{[m]} = (\begin{array}{cccc} B_0 & 0 & \dots & 0 \\ B_1 & B_0 & \ddots & \vdots \\ \vdots & \ddots & \ddots & 0 \\ B_m & \dots & B_1 & B_0 \end{array}). $$ Characteristic elements $C^{[m]}$ assume the same form. Zero curvature representations $A^{[m]},B^{[m]}$ are to be considered under the gauge group consisting of block triangular matrices $$ S^{[m]} = (\begin{array}{cccc} E & 0 & \dots & 0 \\ S_1 & E & \ddots & \vdots \\ \vdots & \ddots & \ddots & 0 \\ S_m & \dots & S_1 & E \end{array}). $$ with unit matrices $E$ in the diagonal positions. By a cohomological argument presented in~\cite[Prop.~1]{spp}, a nontrivial family $A(\lambda),B(\lambda)$ with analytic dependence on $\lambda$ has expansions~\eqref{Tayl} such that $A_1$ or $B_1$ is not zero. Let \eqref{ZCR'}$^{[m]}$,~\eqref{ds}$^{[m]}$ denote the system obtained by substituting $A \to A^{[m]}, B \to B^{[m]}$ into system~\eqref{ZCR'},~\eqref{ds}, for arbitrary $m > 0$. Observe that systems \eqref{ZCR'}$^{[m]}$,~\eqref{ds}$^{[m]}$ are linear in their highest order unknowns $A_m,B_m,C_m$ and can be solved sequentially. Then the applicable cohomological criterion can be summarized as follows. \begin{proposition}[{\cite[Prop.~3]{spp}}] \label{cohom} Let $m > 0$. If $A_1 = B_1 = 0$ for all solutions $A^{[m]},B^{[m]}$ of system~\eqref{ZCR'}$^{[m]}$,~\eqref{ds}$^{[m]}$, then there is no possibility to construct expansions~\eqref{Tayl} of order~$m$, and consequently, the seed zero curvature representation $A_0,B_0$ cannot belong to a nontrivial analytic family. \end{proposition} Finally, to be able to solve system~\eqref{ZCR'}$^{[m]}$,~\eqref{ds}$^{[m]}$, we need to know the normal forms of matrices $A^{[m]},B^{[m]}$. However, the normal forms for $B(\lambda),C(\lambda)$ established above immediately imply the same normal forms for $C_k$ (diagonal) and $B_k$ (symmetric). \section{Results} \label{z:sect} In this section, we present the results of computation of the cohomological obstructions in the case of the nonparametric zero curvature representation~\eqref{AB0} of equation~\eqref{G}. As a sub-result we obtain the first few coefficients $A_k,B_k$ of Taylor expansions~\eqref{Tayl}. As we have seen in the preceding section (Proposition~\ref{cohom}), the problem reduces to solving the system~\eqref{ZCR'}$^{[m]}$,~\eqref{ds}$^{[m]}$ of linear differential equations in total derivatives, for increasing values of~$m$. This is only possible under a suitable restriction on the jet order of the unknowns $A_k,B_k,C_k$, $k > 0$. To start with, we assume dependence on the first-order jets at most. Upon expanding all total derivatives, equations~\eqref{ZCR'}$^{[m]}$,~\eqref{ds}$^{[m]}$ become a large overdetermined system of linear partial differential equations. As such, the system is solvable by computing the passive (or involutive) form under a suitable (elimination) ranking~\cite{R-L-W}. Starting with $m = 1$, we checked that nonzero matrices $A_1,B_1$ depending on second order derivatives exist for all possible determining relations~\eqref{q=F(p)}. When incrementing $m$ to $2$, nontrivial conditions started to appear, but we also reached the boundaries of our available computing resources. Consequently, our present classification results are still incomplete. Nevertheless, we were able to obtain a passive system of differential equations in several cases. Moreover, in two cases we were able to find $A_2,B_2$ explicitly. One of them were the linear Weingarten surfaces~\eqref{lws}. Their integrability is a well established fact~\cite{R-S}, the associated sine-Gordon equation $\phi_{xy} = \sin\phi$ being a textbook example of integrability. The other class emerged as a solution \begin{equation} \label{E} E = \frac{p}{`e^{1 + c/p}}, \qquad c = `const, \end{equation} of the ordinary differential equation $$ \frac{E''}E - (\frac{E'} E)^2 + \frac 2 p \frac{E'}{E} - \frac 1{p^2} = 0. $$ Henceforth we concentrate on the solution~\eqref{E}. The coefficients $u,v,q$ are easily found from~\eqref{E2uvq} to be $$ u = \frac{`e^{1 + c/p}}{p}, \quad v = \frac{p + c}{p `e^{1 + c/p}}, \quad q = \frac{p c}{p + c}. $$ The last equality shows that the condition of constant astigmatism~\eqref{nws1} holds with the constant $-1/c$ on the right-hand side. The Gauss equation~\eqref{G} becomes $$ p_{yy} = \frac{c^2}{`e^{\displayed{4 (1 + \frac cp)}}} p_{xx} + 2 \frac{p + c}{p^2} p_y^2 + 2 \frac{c^2(c - p)}{`e^{\displayed{4 (1 + \frac cp)}} p^2} p_x^2 + \frac{c p^2}{`e^{\displayed{2 (1 + \frac cp)}}}. $$ In principle, the cohomological method we applied can only prove nonintegrability and only indicate, but not prove, integrability. However, it was easy to guess an ansatz based on the form of $A_k$ and $B_k$. By solving~\eqref{ZCR'},~\eqref{ds} we obtained a $\lambda$-dependent zero curvature representation \begin{equation} \label{p:zcr} A = (\begin{array}{cc} \lambda c \frac{p_x}{p^2} + \sqrt{\lambda^2 + \lambda} `e^{\displayed{2 (1 + \frac{c}{p})}} \frac{p_y}{p^2} & \lambda `e^{\displayed{1 + 2 \frac{c}{p}}} \\ (\lambda + 1) `e & -\lambda c \frac{p_x}{p^2} - \sqrt{\lambda^2 + \lambda} `e^{\displayed{2 (1 + \frac{c}{p})}} \frac{p_y}{p^2} \end{array}), \\ B = (\begin{array}{cc} \lambda c \frac{p_y}{p^2} + \sqrt{\lambda^2 + \lambda} c^2 `e^{\displayed{-2 (1 + \frac{c}{p})}} \frac{p_x}{p^2} & \sqrt{\lambda^2 + \lambda} c `e^{-1} \\ \sqrt{\lambda^2 + \lambda} c `e^{\displayed{-1 - 2\frac{c}{p}}} & -\lambda c \frac{p_y}{p^2} - \sqrt{\lambda^2 + \lambda} c^2 `e^{-\displayed{2 (1 + \frac{c}{p})}} \frac{p_x}{p^2} \end{array}), \end{equation} which reduces to the initial $A_0,B_0$ given by~\eqref{AB0} when $\lambda = -\frac12$. The dependence on $p_y$ explains why this class of Weingarten surfaces is missing in Wu's paper~\cite{Wu}. Upon substitution \begin{equation} \label{p2z} x \to \frac x{|c|^{1/4}}, \quad y \to \frac y{|c|^{3/4}}, \quad p \to \frac {4 c}{2 \ln z + \ln |c| - 4} \end{equation} the Gauss equation~\eqref{G} simplifies to $$ \numbered\label{z} z_{yy} + (\frac1z)_{xx} + 2 = 0, $$ and the zero-curvature representation~\eqref{p:zcr} to \begin{equation} \label{z:zcr} A = (\begin{array}{cc} \frac12 \sqrt{\lambda^2 + \lambda} z_y + \frac{1 + 2\lambda}{4} \frac{z_x}{z} & (\lambda + 1) \sqrt z \\ \lambda \sqrt z & -\frac12 \sqrt{\lambda^2 + \lambda} z_y - \frac{1 + 2\lambda}{4} \frac{z_x}{z} \end{array}), \\ B = (\begin{array}{cc} \frac12 \sqrt{\lambda^2 + \lambda} \frac{z_x}{z^2} + \frac{1 + 2\lambda}{4} \frac{z_y}{z} & \frac{\sqrt{\lambda^2 + \lambda}}{\sqrt z} \\ \frac{\sqrt{\lambda^2 + \lambda}}{\sqrt z} & -\frac12 \sqrt{\lambda^2 + \lambda} \frac{z_x}{z^2} - \frac{1 + 2\lambda}{4} \frac{z_y}{z} \end{array}). \end{equation} Let us remark that one can remove the $x$-derivatives from $A$ and $y$-derivatives from $B$ by the gauge transformation~\eqref{gauge}, albeit at the cost of introducing an exponential dependence on the spectral parameter. In~\eqref{p:zcr} and~\eqref{z:zcr}, the corresponding gauge matrix is $$ S = (\begin{array}{cc} `e^{-\frac{\lambda c}{p}} & 0 \\ 0 & `e^{\frac{\lambda c}{p}} \end{array}) \quad\text{and}\quad S = (\begin{array}{cc} z^{\lambda/2} & 0 \\ 0 & z^{-\lambda/2} \end{array}), $$ respectively. For instance, the pair~\eqref{z:zcr} becomes $$ A' = (\begin{array}{cc} \frac12 \sqrt{\lambda^2 + \lambda} z_y & (\lambda + 1) z^{-\lambda} \\ \lambda z^{\lambda + 1} & -\frac12 \sqrt{\lambda^2 + \lambda} z_y \end{array}), \quad B' = (\begin{array}{cc} \left.\frac12\right. \sqrt{\lambda^2 + \lambda} \frac{z_x}{z^2} & \sqrt{\lambda^2 + \lambda} z^{-\lambda - 1} \\ \sqrt{\lambda^2 + \lambda} z^\lambda & -\left.\frac12\right. \sqrt{\lambda^2 + \lambda} \frac{z_x}{z^2} \end{array}). $$ Equation~\eqref{z} has obvious translational symmetries $\partial_x, \partial_y$, the scaling symmetry $2 z \partial_z - x \partial_x + y \partial_y$, and a discrete symmetry $$ \numbered\label{z:ds} x \to y, \quad y \to x, \quad z \to \frac1z. $$ Computation reveals also two third-order symmetries of equation~\eqref{z}. One of them has the generator $$ \numbered\label{z:sym} \padded{\quad} \frac{z^3}{K^3} (z_{xxx} - z z_{xxy}) - \frac3{K^5} z^3 (z_x - z z_y) (z_{xx} - z z_{xy})^2 \\ - \frac2{K^5} z^5 (9 z_x - z z_y) z_{xx} + \frac1{2 K^5} z^2 (9 z_x^2 + 4 z z_x z_y - z^2 z_y^2) (z_x - z z_y) z_{xx} \\ - \frac2{K^5} z^3 z_x (z_x - z z_y) (4 z_x - z z_y) z_{xy} + \frac4{K^5} z^6 z_x z_{xy} \\ + \frac3{K^5} z^4 (5 z_x - z z_y) z_x^2 - \frac3{K^5} z (z_x - z z_y) z_x^4, \return $$ where $$ K = \sqrt{(z_x - z z_y)^2 + 4 z^3}. $$ The other is obtained by conjugation with the discrete symmetry~\eqref{z:ds}. Moreover, A.~Sergyeyev succeeded in finding a recursion operator for equation~\eqref{z} in the usual pseudodifferential form $$ \numbered\label{z:ro} -z_y D_x^{-1} + z_x D_x^{-2} D_y + 2 z D_x^{-1} D_y $$ (unpublished). As far as we could see, the operator generates only nonlocal symmetries. We leave as an open problem to find a recursion operator that would generate the third-order symmetry~\eqref{z:sym}. Let us conclude this section with some easy geometric observations. First of all, we can put $c = 1$ without loss of generality. This can be always achieved by rescaling the ambient Euclidean metric and, if necessary, changing the orientation. Now, the symmetries of the constant astigmatism equation~\eqref{z} have the following geometric interpretation. Translation symmetries are simply reparametrizations of the surface. The scaling symmetry $\phi_\epsilon$: $x \to `e^\epsilon x$, $y \to `e^{-\epsilon} y$, $z \to `e^{-2\epsilon} z$ takes a given surface $\mathbf r(x,y)$ to the parallel surface $\mathbf r(x,y) + \epsilon \mathbf n(x,y)$. This is not surprising since parallel surfaces obviously have equal astigmatism in the corresponding points. Finally, swapping the orientation is another symmetry, which can be identified with a composition of the discrete symmetry~\eqref{z:ds} and the rescaling $\phi_1$. Hence, the discrete symmetry~\eqref{z:ds} corresponds to the change of the orientation followed by taking the parallel surface at the unit distance. \section{Relation to pseudospherical surfaces} As already mentioned in the introduction, nineteenth century geometers knew of a simple relation between pseudospheric surfaces and surfaces of constant astigmatism, even though they did not find the latter important enough to be named. In this section we reproduce some of their findings and derive a nonlocal transformation between the constant astigmatism equation~\eqref{z} and the famous sine-Gordon equation. Again, we put $c = 1$ for simplicity, meaning that the associated focal surfaces will be of Gaussian curvature~$-1$. The forthcoming calculations are conveniently performed in terms of the variable $z$ given by formula~\eqref{p2z} or a new variable $w$ related to $z$ by $$ \numbered\label{z2w} z = `e^{2 w}. $$ Then we have $$ \numbered\label{pquv2w} u = (w - 1) `e^w, \quad v = \frac{w}{`e^w}, \quad p = \frac{1}{w - 1}, \quad q = \frac{1}{w}. $$ and the discrete symmetry~\eqref{z:ds} becomes simply $$ \numbered\label{w:ds} x \to y, \quad y \to x, \quad w \to -w. $$ Given a surface $\mathcal L$, recall that its {\it evolutes} (also known as focal surfaces) are the loci of the principal centres of curvature of $\mathcal L$. Obviously, a generic surface $\mathcal L$ has two evolutes. They interchange positions under the change of the orientation. \begin{proposition}[Ribaucour~\cite{Ri}] \label{prop:Ri} Evolutes of surfaces of constant astigmatism are pseudospherical surfaces. \end{proposition} \begin{proof} Let $\mathbf r(x,y)$ be a surface parametrized by curvature lines. We use the orthonormal frame $(\mathbf e_1,\mathbf e_2,\mathbf n)$, where $$ \mathbf e_1 = \mathbf r_x/u, \quad \mathbf e_2 = \mathbf r_y/v, \quad \mathbf n = \mathbf e_1 \times \mathbf e_2. $$ Then the two evolutes $\mathcal L'$ and $\mathcal L''$ are given by $$ \mathbf r' = \mathbf r + \frac{\mathbf n}p, \qquad \mathbf r'' = \mathbf r + \frac{\mathbf n}q, $$ respectively. An easy calculation using the Gauss--Weingarten formulas~\eqref{G-W:so3} shows that $$ \mathbf r'_x = -\frac{p_x}{p^2} \mathbf n, \qquad \mathbf r'_y = -\frac{p_y}{p^2} \mathbf n + (1 - \frac qp) \mathbf r_y, \\ \mathbf r''_x = -\frac{q_x}{q^2} \mathbf n + (1 - \frac pq) \mathbf r_x, \qquad \mathbf r''_y = -\frac{q_y}{q^2} \mathbf n, $$ the unit normals being $$ \mathbf n' = \frac{\mathbf r_x}u, \qquad \mathbf n'' = \frac{\mathbf r_y}v. $$ Now assume $\mathbf r(x,y)$ to be a surface of constant astigmatism. By applying the substitutions~\eqref{pquv2w} we obtain the first fundamental form of the evolutes in terms of~$w$: $$ `I' = (w_x\,dx + w_y\,dy)^2 + `e^{-2 w}\,dy^2 = dw^2 + `e^{-2 w}\,dy^2, \\ `I'' = `e^{2 w}\,dx^2 + (w_x\,dx + w_y\,dy)^2 = `e^{2 w}\,dx^2 + dw^2. $$ These are the well known pseudospherical metrics in terms of geodesic coordinates $w,y$ and $w,x$ on the first and the second sheet, respectively. \end{proof} For further reference we also compute the second fundamental forms $$ `II' = -`e^w w_x\,dx^2 + \frac{w_x}{`e^{3w}}\,dy^2, \qquad `II'' = `e^{3w} w_y\,dx^2 - \frac{w_y}{`e^w}\,dy^2. $$ Proposition~\ref{prop:Ri} provides as with a couple of transformations from the constant astigmatism equation~\eqref{z} to the sine-Gordon equation. To write them explicitly, we need to equip $\mathcal L'$ and $\mathcal L''$ with the asymptotic coordinates $\xi,\eta$, i.e., the fundamental forms have to be $$ `I' = d\xi^2 + 2\cos \phi' \,d\xi\,d\eta + d\eta^2, \quad `II' = 2\sin \phi' \,d\xi\,d\eta, \\ `I'' = d\xi^2 + 2\cos \phi'' \,d\xi\,d\eta + d\eta^2, \quad `II'' = 2\sin \phi'' \,d\xi\,d\eta. $$ Here $\phi'$ and $\phi''$ are the angles between the coordinate lines on $\mathcal L'$ and $\mathcal L''$, respectively. Using the previous expression of fundamental forms $`I',`II'$ and $`I'',`II''$ in terms of the variable $w$, we easily see that $\xi,\eta$ can be obtained by the ``reciprocal transformation''~\cite{R-S} $$ \numbered\label{xieta} d\xi = \frac12 \sqrt{(w_x + `e^{2 w} w_y)^2 + `e^{2 w}}\,dx + \frac12 \sqrt{(`e^{-2 w} w_x + w_y)^2 + `e^{-2 w}}\,dy, \\ d\eta = \frac12 \sqrt{(w_x - `e^{2 w} w_y)^2 + `e^{2 w}}\,dx - \frac12 \sqrt{(`e^{-2 w} w_x - w_y)^2 + `e^{-2 w}}\,dy. $$ These formulas are valid on both sheets and reflect another property established by Ribaucour~\cite{Ri}, namely that asymptotic lines on $\mathcal L'$ and $\mathcal L''$ correspond. Then the angle $\phi'$ associated with the first sheet satisfies $$ \numbered\label{phi'} \cos\phi' = \frac{w_x^2 - `e^{2w} - `e^{4w} w_y^2} {\sqrt{(w_x + `e^{2 w} w_y)^2 + `e^{2 w}} \sqrt{(w_x - `e^{2 w} w_y)^2 + `e^{2 w}}}, \\ \sin\phi' = -\frac{2 `e^{w} w_x} {\sqrt{(w_x + `e^{2 w} w_y)^2 + `e^{2 w}} \sqrt{(w_x - `e^{2 w} w_y)^2 + `e^{2 w}}}, $$ while the angle $\phi''$ associated with the second sheet satisfies a similar set of equations related by the substitution~\eqref{w:ds}. \begin{proposition} Let $z(x,y)$ be a solution of the constant astigmatism equation~\eqref{z}, let $w = \frac12 \ln z$. Determine function $\phi'$ by formula~\eqref{phi'}, and new coordinates $\xi,\eta$ by the reciprocal transformation~\eqref{xieta}. Then $\phi'(\xi,\eta)$ is a solution of the sine-Gordon equation $\phi_{\xi\eta} = \sin\phi$. \end{proposition} Another solution of the sine-Gordon equation can be obtained by combination with the discrete symmetry~\eqref{w:ds}. The other symmetries (translation and scaling) do not lead to essentially new solutions. Now, it is easy to check that {\it the evolutes of surfaces of constant astigmatism are related by the classical Bianchi transformation.} Indeed, the corresponding points $\mathbf r'$ and $\mathbf r''$ have a constant distance equal to $1/p - 1/q$. The corresponding normals $\mathbf n' = \mathbf r_x/u$ and $\mathbf n'' = \mathbf r_y/v$ are orthogonal. Finally, being directed along the normal vector $\mathbf n$, the line joining the points $\mathbf r'$ and $\mathbf r''$ is tangent to both surfaces $\mathcal L'$ and~$\mathcal L''$. These three properties characterize the classical Bianchi transformation. The Bianchi transformation is, however, superseded by the classical B\"acklund transformation~\cite{BT}, where the condition on the angle between the normals is relaxed from being right to being constant. \section{Surfaces of constant astigmatism as involutes} \label{involut} In principle, all surfaces of constant astigmatism can be obtained from solutions of the sine-Gordon equation as involute surfaces, see, e.g., Darboux~\cite[\S802]{Dar}, Bianchi~\cite[\S130--\S150]{BiaI} or Weatherburn~\cite[Ch.~8]{Wea}. Geodesic nets on pseudospheric surfaces fall into three classes: hyperbolic, parabolic, and elliptic~\cite[\S102]{BiaI}. Of them only the parabolic geodesic nets lead to surfaces of constant astigmatism~\cite[\S136]{BiaI}. Recall that the sine-Gordon $\phi_{\xi\eta} = \sin\phi$ describes surfaces of constant curvature $-1$ in the asymptotic coordinates~$\xi,\eta$. By definition, $$ `I = d\xi^2 + 2 \cos\phi\,d\xi\,d\eta + d\eta^2, \qquad `II = 2 \sin\phi\,d\xi\,d\eta, $$ which leads to the Gauss--Weingarten equations $$ \numbered\label{GW-sG} \mathbf r_{\xi\ax} = \frac{\cos\phi\,\mathbf r_\xi - \mathbf r_\eta}{\sin\phi} \phi_\xi, \quad \mathbf r_{\xi\eta} = \sin\phi\,\mathbf n, \quad \mathbf r_{\eta\ay} = \frac{\cos\phi\,\mathbf r_\eta - \mathbf r_\xi}{\sin\phi} \phi_\eta, \\ \mathbf n_\xi = \frac{\cos\phi\,\mathbf r_\xi - \mathbf r_\eta}{\sin\phi}, \qquad \mathbf n_\eta = \frac{\cos\phi\,\mathbf r_\eta - \mathbf r_\xi}{\sin\phi}. $$ Recall that coordinates $X,Y$ on a pseudospheric surface are called {\it parabolic geodesic} if the first fundamental form can be written as $$ `I = dX^2 + `e^{2X}\,dY^2. $$ To find the transformation from asymptotic to parabolic geodesic coordinates, observe that $d\xi^2 + 2 \cos\phi\,d\xi\,d\eta + d\eta^2 = dX^2 + `e^{2X}\,dY^2$ is equivalent to the system $$ X_\xi^2 + `e^{2X}Y_\xi = 1, \quad X_\xi X_\eta + `e^{2X}Y_\xi Y_\eta = \cos\phi, \quad X_\eta^2 + `e^{2X}Y_\eta = 1. $$ This system can be rewritten as $$ \numbered\label{g2a} X_\xi = \cos\alpha, \qquad\qquad Y_\xi = `e^{-X} \sin\alpha, \\ X_\eta = \cos\beta, \qquad\qquad Y_\eta = `e^{-X} \sin\beta, $$ and $$ \numbered\label{pab} \phi = \alpha - \beta. $$ In fact,~\eqref{pab} could be also $\phi = \beta - \alpha$, which can be reversed by changing the orientation of the surface. The new unknowns $\alpha$ and $\beta$ can be identified with the angles between the geodesics and the two asymptotic coordinate lines. The integrability conditions of system~\eqref{g2a} are $$ \numbered\label{ab} \beta_\xi = -\sin \alpha, \qquad \alpha_\eta = -\sin\beta, $$ or, in view of relation~\eqref{pab}, $$ \numbered\label{b2a} \beta_\xi = -\sin(\phi + \beta), \qquad \beta_\eta = -\phi_\eta - \sin\beta. $$ These are already compatible by virtue of the sine-Gordon equation. From equations~\eqref{g2a} we obtain $$ \mathbf r_X = -\frac{\sin\beta}{\sin\phi} \mathbf r_\xi + \frac{\sin\alpha}{\sin\phi} \mathbf r_\eta, \qquad \mathbf r_Y = (\frac{\cos\beta}{\sin\phi} \mathbf r_\xi + \frac{\cos\alpha}{\sin\phi} \mathbf r_\eta)`e^X. $$ With respect to a given geodesic net, the involute surface $\tilde{\mathbf r}$ is composed of individual involute curves to the geodesics, based on one and the same orthogonal line $Y = `const$. Hence, $$ \tilde{\mathbf r} = \mathbf r + (a - X) \mathbf r_X, $$ where $a$ is an arbitrary constant. With the help of equations~\eqref{GW-sG}, the fundamental forms $\tilde{`I},\tilde{`II}$ of the involute surface $\tilde{\mathbf r}$ can be routinely computed in asymptotic coordinates. In particular, the unit normal is $\tilde{\mathbf n} = \mathbf r_X$ and $$ \tilde{`I} = (X^2 - X + \frac12) (1 - \cos 2\alpha)\,d\xi^2 + (2X - 1) (\cos(\alpha + \beta) - \cos\phi)\,d\xi\,d\eta \\\quad + (X^2 - X + \frac12) (1 - \cos 2\beta)\,d\eta^2, \\ \tilde{`II} = (X - \frac12) (\cos 2\alpha - 1)\,d\xi^2 + (\cos(\alpha + \beta) - \cos\phi)\,d\xi\,d\eta \\\quad + (X - \frac12) (\cos 2\beta - 1)\,d\eta^2. $$ Hence, the principal radii of curvature are $X$, $X - 1$. The Gauss--Mainardi--Codazzi equations of the involute surface hold as a consequence of the sine--Gordon equation, the two equations~\eqref{g2a} on $X$ and the system~\eqref{b2a} on~$\beta$. To obtain the corresponding solution of the constant astigmatism equation~\eqref{z}, we have to reparametrize the involute surfaces by curvature lines. Let $x,y$ denote the new coordinates. We choose $x = Y$ and define $y$ by the compatible system of equations $$ \numbered\label{y2a} y_\xi = `e^X \sin\alpha, \qquad y_\eta = `e^X \sin\beta. $$ A routine calculation shows that $`e^{-2X(x,y)}$ is a solution of the constant astigmatism equation~\eqref{z}. Summarizing, we have the following proposition. \begin{proposition} Let $\phi(\xi,\eta)$ be a solution of the sine-Gordon equation $\phi_{\xi\eta} = \sin\phi$. Let $\alpha,\beta$ be solutions of the compatible equations $$ \beta_\xi = -\sin \alpha, \qquad \alpha_\eta = -\sin\beta, \qquad \alpha - \beta = \phi. $$ Determine functions $X,x,y$ by equations $$ dX = \cos\alpha\,d\xi + \cos\beta\,d\eta, \\ dx = `e^{-X} (\sin\alpha\,d\xi + \sin\beta\,d\eta), \\ dy = `e^X (\sin\alpha\,d\xi + \sin\beta\,d\eta). $$ Then the function $`e^{-2X(x,y)}$ is a solution of the constant astigmatism equation~\eqref{z}. \end{proposition} \begin{example} \rm {\it Von Lilienthal's surfaces} (involutes of the pseudosphere). Published in 1887, these surfaces seem to have fallen into oblivion. Recall that the pseudosphere is a surface obtained by rotating the tractrix around its asymptote. The meridians are geodesics of the parabolic type and therefore von Lilienthal's surface is obtained by rotating the involute of the tractrix (which itself is the involute of the catenary). In geodesic coordinates $X,Y$, the ``upper half'' of the pseudosphere has a parametrization $$ \mathbf r = (\begin{array}{c} `e^{-X} \cos Y \\ `e^{-X} \sin Y \\ `arcosh`e^{X} - \sqrt{1 - `e^{-2 X}} \end{array}), \quad X > 0, $$ whose first fundamental form is $dX^2 + `e^{-2 X}\,dY^2$ (differs by the sign of $X$ from the canonical form used in the preceding section). Then $$ \tilde{\mathbf r} = \mathbf r + (a - X) \mathbf r_X = (\begin{array}{c} (X - a + 1) `e^{-X} \cos Y \\ (X - a + 1) `e^{-X} \sin Y \\ `arcosh`e^X - (X - a + 1) \sqrt{1 - `e^{-2 X}} \end{array}), \quad X > 0, $$ parametrizes a rotational surface, for every real constant $a$. The surface is regular for all $a \le 0$. Otherwise it has a cuspidal edge at $X = a$, which is a circle of radius $`e^{-a}$. Another singularity that occurs for every $a > 1$ is the intersection with the rotation axis at $X = a - 1$. Choosing the orientation so that the normal vector is $$ \tilde{\mathbf n} = (\begin{array}{c} -`e^{-X} \cos Y \\ -`e^{-X} \sin Y \\ \sqrt{1 - `e^{-2 X}} \end{array}) $$ (i.e., $\mathbf n$ swaps orientation when crossing either of the singularities), then $$ \tilde{`I} = \frac{(X - a)^2}{`e^{2 X} - 1}\,dX^2 + \frac{(X - a + 1)^2}{`e^{2 X}}\,dY^2, \\ \tilde{`II} = \frac{X - a}{`e^{2 X} - 1}\,dX^2 + \frac{X - a + 1}{`e^{2 X}}\,dY^2. $$ and the principal radii of curvature are $X - a$ and $X - a + 1$. The corresponding solution of the constant astigmatism equation~\eqref{z} is $$ z = \frac1{x^2 - `e^{2(a - 1)}}. $$ \setlength{\unitlength}{1.4in} \newcommand{\obr}[3] { \begin{picture}(1,1)(0,0) \put(0.6,0.5){\makebox(0,0){\includegraphics[width=3in]{#2}}} \put(0.6,0.5){\makebox(0,0){\vrule depth 0cm height 1.6in width .05mm}} \put(0.6,0.5){\makebox(0,0){\includegraphics[width=3in]{#1}}} \put(0.6,0){\makebox(0,-0.35){\small $a = #3$}} \end{picture} } \begin{figure}[h] \vskip 1pc \begin{center} \obr{plot_1.ps}{plot_common.ps}{-0.5} \obr{plot_2.ps}{plot_common.ps}{0} \obr{plot_3.ps}{plot_common.ps}{0.2} \obr{plot_4.ps}{plot_common.ps}{1-\ln 2} \vspace{0.7in} \obr{plot_5.ps}{plot_common.ps}{0.7} \obr{plot_6.ps}{plot_common.ps}{1} \obr{plot_7.ps}{plot_common.ps}{1.2} \obr{plot_8.ps}{plot_common.ps}{1.7} \end{center} \vskip 1pc \caption{A gallery of von Lilienthal surfaces} \label{Pictures} \end{figure} Plane sections of von Lilienthal surfaces for various values of the parameter $a$ can be seen on Figure~\ref{Pictures}. Besides the rotation axis, each picture shows the tractrix, which is the plane section of the pseudosphere, and its involute curve, which is the plane section of the von Lilienthal surface. We finish this example with a short exploration of the behaviour at the limits of the definition domain. For $X = \infty$ the surface closes up at a point on the rotation axis at the height $a - 1 + \ln 2$, where both principal radii of curvature are infinite (the zero height is that of the cusp of the tractrix). For $X \to 0$ the surface vertically approaches a horizontal circle of diameter $|1 - a|$. Two surfaces $\tilde{\mathbf r}(X,Y)$ and $-\tilde{\mathbf r}(X,Y)$ can be glued along this circle to form a single surface of constant astigmatism~$1$. For $a = 1$ both glued surfaces have a cusp here. \end{example} \section{Conclusions and discussion} Among the still incomplete results of classification of integrable Weingarten surfaces, we have identified a class originally introduced and investigated by nineteenth-century geometers. The class, which we propose to call surfaces of constant astigmatism, is governed by the equation $$ z_{yy} + (\frac1z)_{xx} + 2 = 0. $$ For this equation we found an $\mathfrak{sl}(2)$-valued zero curvature representation depending on a parameter, a third-order symmetry, and a nonlocal transformation to the sine-Gordon equation $\phi_{\xi\eta} = \sin\phi$. We had to leave aside the problem of finding a B\"acklund transformation as well as a recursion operator producing a hierarchy of local symmetries. It should be stressed that the classification problem of integrable surfaces is far from being easy. An obvious reason lies in the abundance of integrability-preserving ways to derive one surface from another. Clearly, parallel surfaces, evolutes, and involutes of integrable surfaces are integrable. On the differential equation level, the corresponding notion is that of the covering~\cite{K-V}. The integrable classes of surfaces must be either closed with respect to taking derived surfaces or the derivation must map one integrable class into another. \ack This paper would be impossible without encouragement, support and advice from J.~Cie\'sli\'nski, E. Ferapontov, R. L\'opez and A. Sergyeyev. The first-named author was supported by GA\v{C}R under project 201/07/P224. The second-named author by M\v{S}MT under project MSM~4781305904. Thanks are also due to CESNET for granting access to the MetaCentrum computing facilities. \section*{References}
\section{Results and Discussion} Figure~\ref{fig:jet} shows the resulting per-trigger jet pair yields for selected trigger-partner combinations in $p\mathord{+}p$ and the 20\% most central Au$\mathord{+}$Au collisions. \begin{figure}[tb] \centering \includegraphics[width=1.0\linewidth]{PTYsix.eps} \caption{\label{fig:jet} (Color online) Per-trigger jet pair yield vs. $\Delta\phi$ for selected \mbox{$\pi^{0}$}\xspace trigger and $h^\pm$ partner $p_T$ combinations (\mbox{$p_T^t$}\xspace $\otimes$ \mbox{$p_T^a$}\xspace) in Au$\mathord{+}$Au and $p\mathord{+}p$ collisions (solid and open symbols, respectively). Depicted Au$\mathord{+}$Au systematic uncertainties include point-to-point correlated background level and modulation uncertainties (gray bands and open boxes, respectively). For shape comparison insets show away-side distributions scaled to match at $\Delta\phi = \pi$. } \end{figure} On the near side, the widths in central Au$\mathord{+}$Au are comparable to $p\mathord{+}p$ over the full \mbox{$p_T^t$}\xspace and \mbox{$p_T^a$}\xspace ranges, while the yields are slightly enhanced at low \mbox{$p_T$}\xspace, matching $p\mathord{+}p$ as \mbox{$p_T$}\xspace increases. On the opposing side, qualitatively one observes that for low \mbox{$p_T^t$}\xspace and low \mbox{$p_T^a$}\xspace the Au$\mathord{+}$Au jet peaks are strongly broadened and non-Gaussian. In contrast, at high \mbox{$p_T^t$}\xspace and high \mbox{$p_T^a$}\xspace the yield is substantially suppressed, but the shape appears consistent with the measurement in the $p\mathord{+}p$ case (as has been previously reported in much broader \mbox{$p_T$}\xspace ranges for unidentified charged hadron triggers~\cite{ppg083,starhighpt}). Here we quantify the trends in the shape and yield between these two extremes. First, we have performed a fit to the away-side distribution over the range $|\Delta \phi - \pi| < \pi/2$ to a simple Gaussian distribution. Figure~\ref{fig:sigma} shows the results. In $p\mathord{+}p$ collisions, the away-side width narrows at higher trigger and partner momentum as expected from the angular ordering of jet fragmentation. For \mbox{$p_T^t$}\xspace $>$ 7 GeV/$c$, the widths are consistent within uncertainties between $p\mathord{+}p$ and Au$\mathord{+}$Au at all \mbox{$p_T^a$}\xspace. There is no evidence of large jet broadening from in-medium scattering~\cite{deflect} or from initial state effects~\cite{heavyIonKT}, expected for surviving partons produced in the interior rather than the surface of the medium. However, it is also possible that for high \mbox{$p_T^t$}\xspace the broadening is modest for the leading parton and its fragmentation products and the radiated energy results in only very low \mbox{$p_T^a$}\xspace hadrons (mostly with \mbox{$p_T^a$}\xspace $<$ 0.5 GeV/c). For \mbox{$p_T^t$}\xspace $<$ 7 GeV/$c$, the away-side widths are significantly wider than in $p\mathord{+}p$, except at the highest \mbox{$p_T^a$}\xspace. Note that for \mbox{$p_T^t$}\xspace $<$ 7 GeV/$c$ and low \mbox{$p_T^a$}\xspace, the best fit $\sigma_{\rm away}$ values are larger than $\pi/2$ radians. These trends in shape are further quantified with the use of a $\chi^2$ test to examine the hypothesis that the central Au$\mathord{+}$Au jet shape on the near and away side is the same as the $p\mathord{+}p$ jet shape. For \mbox{$p_T^t$}\xspace $>$7 GeV/$c$, agreement is found for all \mbox{$p_T^a$}\xspace. However, for \mbox{$p_T^t$}\xspace at 5--7 (4--5) GeV/$c$, the agreement worsens sharply for \mbox{$p_T^a$}\xspace $<$ 3 (4) GeV/$c$ as the away-side jet becomes increasingly broad. For example, the p-values for agreement between the $p\mathord{+}p$ and Au$\mathord{+}$Au shapes for \mbox{$p_T^a$}\xspace = 1-2 GeV/$c$ are very small ($<$ $10^{-4}$) for \mbox{$p_T^t$}\xspace = 4--5 and 5--7 GeV/$c$, but indicate reasonable agreement (0.33 and 0.16) for \mbox{$p_T^t$}\xspace = 7--9 and 9--12 GeV/$c$, respectively. The statistical precision of the experimental data does not allow conclusion of a sharp transition in the shape; however, there is a clear indication of a trend towards either much smaller modification or unmodifed jet shapes for higher \mbox{$p_T^t$}\xspace at all \mbox{$p_T^a$}\xspace. To confirm this finding, we compared the away-side distributions in Au$\mathord{+}$Au central events for \mbox{$p_T^t$}\xspace 5--7 GeV/$c$ with \mbox{$p_T^t$}\xspace 7--9 GeV/$c$ for \mbox{$p_T^a$}\xspace 1--2 GeV/$c$ (see Fig. 1) and find the probability that they have a common source is small (p-value $<$ 0.07). \begin{figure}[tb] \centering \includegraphics[width=1.0\linewidth]{sigmaAWAY.eps} \caption{\label{fig:sigma} (Color online) Away-side jet widths from a Gaussian fit by $h^\pm$ partner momentum for various \mbox{$\pi^{0}$}\xspace trigger momenta in $p\mathord{+}p$ (open circles), midcentral 20--60\% Au$\mathord{+}$Au (solid circles), and central 0--20\% Au$\mathord{+}$Au collisions (squares). For comparison, an interpolation of the $p\mathord{+}p$ is depicted (curve). In cases where the best fit $\sigma_{\rm away} > \pi/2$ radians, the point is off the plot. } \end{figure} The lack of large away-side shape modification for \mbox{$p_T^t$}\xspace $>$ 7 GeV/$c$ and \mbox{$p_T^a$}\xspace $<$ 3 GeV/$c$ is surprising as medium response effects are not generally expected to decrease at larger \mbox{$p_T^t$}\xspace. In descriptions where the medium-induced energy loss ($\Delta E$) is nearly proportional to the initial parton energy ($E$)~\cite{Kharzeev:2008he}, and where the lost energy produces a medium response, a larger medium modification is expected for higher momentum partons. Within our statistical precision, no evidence for this is seen; rather, the opposite is found. However, should $\Delta E/E$ fall steeply with increasing parton \mbox{$p_T$}\xspace, an increased contribution from partons which have lost little energy could make an observation of the medium response more difficult. In alternative models of fluctuating background correlations~\cite{ps,Takahashi:2009na}, the modification is predicted to diminish at higher trigger \mbox{$p_T$}\xspace as the background contribution drops, in agreement with observations. In addition to the shape modification measurement, the away-side integrated yield is determined. Away-side jet yield modification in central collisions, shown in Fig.~\ref{fig:iaa}, is measured by $I_{\rm AA}$ (the ratio of conditional jet pair yields integrated over a particular range in $\Delta \phi$ in Au$\mathord{+}$Au to $p\mathord{+}p$). The $I_{\rm AA}$ uncertainties include uncorrelated errors ($\sigma_{\rm stat}$), point-to-point correlated errors from the background subtraction ($\sigma_{\rm syst}$), and a normalization uncertainty from the single particle efficiency determination. Away-side $I_{\rm AA}$ values for \mbox{$p_T^t$}\xspace $>$ 7 GeV/$c$ tend to fall with \mbox{$p_T^a$}\xspace for both the full away-side region ($|\Delta\phi-\pi| < \pi/2$) and for a narrower ``head'' selection ($|\Delta\phi-\pi| < \pi/6$) until \mbox{$p_T^a$}\xspace $\approx$ 2--3 GeV/c, above which they become roughly constant. The yield enhancement at \mbox{$p_T^t$}\xspace $>$ 7 GeV/$c$ and \mbox{$p_T^a$}\xspace $<$ 2 GeV/$c$ is modest and occurs without significant shape modification (Fig.~\ref{fig:sigma}). When \mbox{$p_T^t$}\xspace is decreased, the away-side $I_{\rm AA}$ differs between the two angular selections as the shape becomes modified. \begin{figure}[tb] \centering \includegraphics[width=1.0\linewidth]{iaaHeadAway.eps} \caption{\label{fig:iaa} (Color online) Away-side $I_{\rm AA}$ for a narrow ``head'' $|\Delta\phi-\pi|<\pi/6$ selection (solid squares) and the entire away-side, $|\Delta\phi-\pi|<\pi/2$ (solid circles) vs. $h^\pm$ partner momentum for various \mbox{$\pi^{0}$}\xspace trigger momenta. Calculations from two different predictions are shown for the head region in applicable \mbox{$p_T$}\xspace ranges. A point-to-point uncorrelated 6\% normalization uncertainty (mainly due to efficiency corrections) applies to all measurements. For comparison, \mbox{$\pi^{0}$}\xspace \mbox{$R_{\rm AA}$}\xspace \cite{Adare:2008qa} bands are included where \mbox{$p_T^t$}\xspace $>$ 5 GeV/$c$. } \end{figure} Average away-side $I_{\rm AA}$ values from weighted averages of the ``head'' region data in Fig.~\ref{fig:iaa} for $\mbox{$p_T^t$}\xspace(\mbox{$p_T^a$}\xspace) > 5(2)$ GeV/$c$ are listed in Table~\ref{tab:iaa} for both central and midcentral collisions. The fits, which are not shown, cover the momentum range where shape modification is weak or nonexistent. \begin{table}[tb] \caption{\label{tab:iaa} Average away-side $I^{\rm head}_{\rm AA}$ above 2 GeV/$c$ for various \mbox{$\pi^{0}$}\xspace trigger momenta in central and midcentral collisions where $|\Delta\phi-\pi|<\pi/6$. Note: a 6\% scale uncertainty applies to all $I_{\rm AA}$ values. } \begin{ruledtabular} \begin{tabular}{ccccccc} \multicolumn{1}{c}{ } & \multicolumn{3}{c}{Cent 0--20\%} & \multicolumn{3}{c}{Cent 20--60\%} \\ $p^{t}_{T}$ & $I^{\rm head}_{\rm AA}$ & $\pm\sigma_{\rm stat}$ & $\pm\sigma_{\rm syst}$ & $I^{\rm head}_{\rm AA}$ & $\pm\sigma_{\rm stat}$ & $\pm\sigma_{\rm syst}$ \\ \hline 5--7 & 0.35 & 0.04 & 0.03 & 0.55 & 0.02 & 0.04 \\ 7--9 & 0.34 & 0.05 & 0.03 & 0.64 & 0.04 & 0.02 \\ 9--12 & 0.50 & 0.08 & 0.02 & 0.73 & 0.06 & 0.02 \\ \end{tabular} \end{ruledtabular} \end{table} The away-side $I_{\rm AA}$ values for both centrality selections tend to rise as \mbox{$p_T^t$}\xspace increases. Reference~\cite{starhighpt} measured a constant away-side $I_{\rm AA}$ for $z_T$ ($=p^{a}_T/p^{t}_T$) above 0.4 for triggers at 8--16 GeV/$c$, but such a single point spanning a broad momentum range fails to provide information on the \mbox{$p_T^t$}\xspace evolution of \mbox{$I_{\rm AA}$}\xspace for comparison with the present results. Figure~\ref{fig:iaa} also shows the \mbox{$\pi^{0}$}\xspace \mbox{$R_{\rm AA}$}\xspace for \mbox{$p_T$}\xspace $>$ 5 GeV/$c$~\cite{Adare:2008qa}. The comparison reveals that \mbox{$I_{\rm AA}$}\xspace is consistently higher than \mbox{$R_{\rm AA}$}\xspace. This feature probably results from a few competing effects. Selection of a high \mbox{$p_T$}\xspace trigger \mbox{$\pi^{0}$}\xspace is expected to bias the hard scattering towards the medium surface. Thus, away-side partons have a long average path length through the medium and consequently lose more energy. However, this does not require that \mbox{$I_{\rm AA}$}\xspace be lower than \mbox{$R_{\rm AA}$}\xspace. The away-side conditional spectrum falls less steeply than the inclusive hadron spectrum and so the same spectral shift will more strongly reduce \mbox{$R_{\rm AA}$}\xspace. Figure~\ref{fig:iaa} also shows \mbox{$I_{\rm AA}$}\xspace calculations from the ACHNS~\cite{Armesto:2009zi} and ZOWW~\cite{Zhang:2009rn} models. Each calculation includes the combination of a parton energy loss formalism and a modeling of medium geometry. The ACHNS calculation, which employs a hydrodynamic evolution model of the medium and an energy loss prescription based on quenching parameters constrained by other data~\cite{Adare:2008qa,starhighpt}, predicts \mbox{$I_{\rm AA}$}\xspace $\lesssim$ \mbox{$R_{\rm AA}$}\xspace. The ZOWW calculation, which utilizes a simple spherical nuclear geometry and is similarly constrained by other data~\cite{Adare:2008qa,starhighpt}, predicts \mbox{$I_{\rm AA}$}\xspace $>$ \mbox{$R_{\rm AA}$}\xspace in agreement with these data. It would be instructive to re-calculate these \mbox{$I_{\rm AA}$}\xspace predictions with a common medium geometry (as was done for \mbox{$R_{\rm AA}$}\xspace in Reference~\cite{Bass:2008rv}) to disentangle the model differences. Additionally, a full assessment including all \mbox{$R_{\rm AA}$}\xspace and \mbox{$I_{\rm AA}$}\xspace measurements, including direct photon trigger data~\cite{ppg090,stardirectph}, is warranted. In summary, \mbox{$\pi^{0}$}\xspace-\mbox{$h^{\pm}$}\xspace correlations over a very broad range in trigger and partner \mbox{$p_T$}\xspace have been measured. We observe an away-side modification for moderate \mbox{$p_T$}\xspace triggers (\mbox{$p_T^t$}\xspace $<$ 7 GeV/$c$) and low \mbox{$p_T$}\xspace partners (\mbox{$p_T^a$}\xspace $<$ 3 GeV/$c$) as has been observed in unidentified dihadron correlations. However, this modification is reduced or absent for triggers above 7 GeV/$c$ for any partner \mbox{$p_T$}\xspace and challenges descriptions where more (initially) energetic partons lose more energy and should produce a larger medium response. At large momenta, i.e. triggers above 5 GeV/$c$ and partners above 2 GeV/$c$, away-side modification factor \mbox{$I_{\rm AA}$}\xspace is above the inclusive \mbox{$\pi^{0}$}\xspace modification factor \mbox{$R_{\rm AA}$}\xspace(\mbox{$p_T$}\xspace$>$ 5 GeV/$c$). We thank the staff of the Collider-Accelerator and Physics Departments at BNL for their vital contributions. We acknowledge support from the Office of Nuclear Physics in DOE Office of Science, NSF, and a sponsored research grant from Renaissance Technologies (USA), MEXT and JSPS (Japan), CNPq and FAPESP (Brazil), NSFC (China), MSMT (Czech Republic), IN2P3/CNRS and CEA (France), BMBF, DAAD, and AvH (Germany), OTKA (Hungary), DAE and DST (India), ISF (Israel), NRF (Korea), MES, RAS, and FAAE (Russia), VR and KAW (Sweden), U.S. CRDF for the FSU, US-Hungary Fulbright, and US-Israel BSF.
\section{Introduction} \label{intro} A deep understanding of collective phenomena in Statistical Mechanics is well established in terms of microscopic spin models. Useful macroscopic descriptions of these models in terms of mean field approaches, pair and higher order approximations, and field theories are also well known. Partly inspired by this success, collective social phenomena are being currently studied in terms of microscopic models of interacting agents \cite{Castellano-09-a}. Agents, playing here the role of spins, sit in the nodes of a network of social interactions and change their state (social option) according to specified dynamical rules of interaction with their neighbors in the network. A general question addressed is the consensus problem, reminiscent of order-disorder transitions: The aim is to establish ranges of the parameters determining the interaction rules and network characteristics for which the system is eventually dominated by one state or option or, on the contrary, when a configuration of global coexistence is reached \cite{SanMiguel-05_CISE}. Language competition falls within the context of such social consensus problems: One considers a population of agents that can use either of two languages (two states). The agents change their state of using one or the other language, by interactions with other agents. One is here interested in determining when a state of dominance (or extinction) of one language is reached, or when a state of global language coexistence, with a finite fraction of the two kind of speakers, prevails. A particular and interesting ingredient in this problem is the possibility of a third state associated with bilingual agents, which have been claimed to play an essential role in processes of language contact \cite{Appel_1987,Mira-05}. A good deal of work along these lines originates in a paper by Abrams and Strogatz \cite{Abrams-03}. These authors introduced a simple population dynamics model with the aim of describing the irreversible death of many languages around the world \cite{Crystal-00}. The original Abrams-Strogatz model (ASM) was a macroscopic description based on ordinary differential equations, but a corresponding microscopic agent based model was described in \cite{Stauffer-07}. This model features probabilities to switch languages determined by the local density of speakers of the opposite language, and by two parameters that we call \emph{prestige}, $\mathcal S$ and \emph{volatility}, $a$. Prestige is a symmetry breaking parameter favoring the state associated with one or other language which accounts for the differences in the social status between the two languages in competition. The volatility parameter determines the functional form of the switching probability. It characterizes a property of the social dynamics associated to the inertia of an agent to change its current option, with its neutral value $a=1$ corresponding to a mechanism of random imitation of a neighbor in the network. Other studies of the ASM account for the effects of geographical boundaries introduced in terms of reaction-diffusion equations \cite{Patriarca_2008} or for Lotka-Volterra modifications of the original model \cite{Pinasco-06}. Another different class of models accounts for many languages in competition, with the aim of reproducing the distribution of language sizes in the world \cite{Schulze-05,Vivane-06(1)}. While in the original paper of Abrams and Strogatz \cite{Abrams-03} the two parameters $\mathcal S$ and $a$ were fitted to a particular linguistic data, most subsequent work has focused on theoretical analysis for the case of symmetrical prestige and neutral volatility. For these parameter values (symmetric $\mathcal S$ and $a=1$) the microscopic ASM coincides with the voter model \cite{Ligget_1985}, a paradigmatic spin model of nonequilibrium dynamics \cite{Marro_Dickman_1999}. Inspired in the modifications proposed by Minett and Wang of the ASM \cite{Wang-05_TRENDS_Ecology,Minett-08}, a microscopic Bilinguals model (BM) which introduces a third (intermediate) state to account for bilingualism has been studied for the case of symmetric $\mathcal S$ and $a=1$ in \cite{Castello-06}. In this way, this case is an extension of the original voter model. The emphasis has been in describing the effects of the third state of bilingual agents in the dynamics of language competition as compared with the reference case provided by the voter model. This includes the characterization of the different processes of domain growth \cite{Castello-06}, and the role of the network topology, like small world networks \cite{Castello-06} and networks with mesoscopic community structure \cite{Castello-07,Toivonen-08}. Other studies associated to variations of the voter model dynamics and the addition of intermediate states have also been addressed in \cite{Dall'Asta-07,Stark-08,blythe2009gmc,Vazquez-03,Vazquez-04,Dall'Asta-08}. A pending task in the study of this class of models for language competition is, therefore, the detailed analysis of the role of the prestige and volatility in their general dynamical properties. In addition, for the voter model, macroscopic field theory descriptions \cite{Al_Hammal_Chate-05,Vazquez-08c} as well as macroscopic and analytical solutions in different complex networks \cite{Vazquez-08a} have been reported, but there is still a lack of useful macroscopic descriptions of these models for arbitrary values of the prestige and volatility parameters. The general aim of this paper is then, to explore the behavior of these models for a wide range of these parameters values, and to derive appropriate macroscopic descriptions that account for the observed order-disorder (language dominance-coexistence) transitions in the volatility-prestige parameter space. In particular, we analyze how the introduction of an intermediate bilingual state affects language coexistence, by comparing the regions of coexistence and one-language dominance of the ASM and BM in the parameter space. In addition, we study how these regions are modified, within the same models, when the dynamics takes place on networks with different topologies. The paper is organized as follows. In section \ref{AS-model}, we introduce and study the Abrams-Strogatz model on fully connected and complex networks. Starting from the microscopic dynamics, we derive ordinary differential equations for the global magnetization (difference between the fraction of speakers of each language) and the interface density (fraction of links connecting opposite-language speakers). We use these equations to analyze ordering and stability properties of the system, that is, whether there is language coexistence or monolingual dominance in the long time limit. In section \ref{MW-model}, we introduce and study the Bilinguals model, following an approach similar to the one in section \ref{AS-model}. In section \ref{Square}, we address the behavior of these language competition models and the order-disorder transitions on square lattices. In particular, we build a macroscopic description of the dynamics of the ASM on square lattices by deriving partial differential equations for the magnetization field, that depend on space and time. Finally, in section \ref{Summary} we present a discussion and a summary of the results. \section{Abrams and Strogatz model} \label{AS-model} The microscopic agent based version \cite{Stauffer-07,Castello-06} of the model proposed by Abrams and Strogatz \cite{Abrams-03} considers a population of N individuals sitting in the nodes of a social network of interactions. Every individual can speak two languages $X$ and $Y$, but it uses only one at a time. In a time step, an individual chosen at random is given the possibility to give up the use of its language and start using the other language. The likelihood that the individual changes language use depends on the fraction of its neighbors using the opposite language. Neighbors are here understood as agents sitting in nodes directly connected by a link of the network. The language switching probabilities are defined as \begin{eqnarray} P(X \to Y) &=& (1-\mathcal S) \, \sigma_y^a ~~~\mbox{and} \nonumber \\ P(Y \to X) &=& \mathcal S \, \sigma_x^a, \label{PXY} \end{eqnarray} where $\sigma_x (\sigma_y)$ is the density of $X (Y)$ neighboring speakers of a given individual, $0 \le \mathcal S \le 1$ is the \emph{prestige} of language $X$, and $a > 0$ is the \emph{volatility} parameter. $\mathcal S$ controls the asymmetry of language change [$\mathcal S>1/2~(\mathcal S <1/2)$ favoring language $X (Y)$], whereas $a$ measures the tendency to switch language use. The case a=1 is a neutral situation, in which the transition probabilities depend linearly on the local densities. A high volatility regime regime exists for $a < 1$, with a probability of changing language state above the neutral case, and therefore agents change their state rather frequently. A low volatility regime exists for $a > 1$ with a probability of changing language state below the neutral case with agents having a larger resistance or inertia to change their state. Having defined the model, we would like to investigate the dynamics and ultimate fate of the population, that is, whether all individuals will agree after many interactions in the use of one language or not. In order to perform an analytical and numerical study of the evolution of the system we consider, in an analogy to spin models, $X$ and $Y$ speakers as spin particles in states $s=-1$ (spin down) and $s=1$ (spin up) respectively. Thus, the state of the system in a given time can be characterized quite well by two macroscopic quantities: The global magnetization $m \equiv \frac{1}{N} \sum_{i=1}^N S_i$, ($S_i$ with $i=1,..,N$ is the state of individual $i$ in a population of size $N$), and the density of pairs of neighbors using different languages $\rho \equiv \frac{1}{2 N_l}\sum_{<i j>} (1-S_i S_j)/2$, where $N_l$ is the number of links in the network and the sum is over all pair of neighbors. The magnetization measures the balance in the fractions of $X$ and $Y$ speakers ($m=0$ corresponding to the perfectly balanced case), whereas $\rho$ measures the degree of disorder in the system. The case $|m|=1$ and $\rho = 0$ corresponds to the totally ordered situation, with all individuals using the same language, while $|m|<1$ and $\rho >0$ indicates that the system is disordered, composed by both type of speakers. The aim is to obtain differential equations for the time evolution of the average values of $m$ and $\rho$. These equations are useful in the study of the properties of the system, from an analytical point of view. We start by deriving these equations in the case of a highly connected society with no social structure (fully connected network), that corresponds to the simplified assumption of a ``well mixed'' population, widely used in population dynamics. We then obtain the equations in a more realistic scenario, when the topology of interactions between individuals is a social complex network. We shall see that the results depend on the particular properties of the network under consideration, reflected in the moments of the degree distribution. \subsection{Fully connected networks} \label{Fully} We consider a network composed by $N$ nodes, in which each node has a connection to any other node. In a time step $\delta t = 1/N$, a node $i$ with state $s$ ($s=\pm 1$) is chosen with probability $\sigma_s$. Then, according to the transitions (\ref{PXY}), $i$ switches its sate with probability \begin{equation} P(s \to -s) = \frac{1}{2}(1-s v)\left( \sigma_{-s} \right)^a, \label{Ps-s} \end{equation} where $\sigma_{-s}$ is the density of neighbors of $i$ with state $-s$, that in this fully connected network is equal to the global density of $-s$ nodes. Given that the total number of individuals is conserved we have that $\sigma_- + \sigma_+ =1$. We define the bias $v \equiv 1-2\mathcal S$ ($-1<v<1$) as a measure of the preference for one of the two languages, with $v>0$ ($v<0$) favoring the $s=1$ ($s=-1$) state. In the case that the switch occurs, the density $\sigma_s$ is reduced by $1/N$, for which the magnetization $m=\sigma_+ - \sigma_-$ changes by $-2 s/N$. Then, the average change in the magnetization can be written as \begin{equation} \frac{dm(t)}{dt} = \frac{1}{1/N} \left[ \sigma_- P(- \to +) \frac{2}{N} - \sigma_+ P(+ \to -) \frac{2}{N} \right]. \label{dmdt0} \end{equation} Using Eq.~(\ref{Ps-s}) and expressing the global densities $\sigma_{\pm}$ in terms of the magnetization, $\sigma_{\pm}=(1 \pm m)/2$, we arrive to \begin{equation} \frac{dm(t)}{dt} = 2^{-(a+1)} (1-m^2) \left[ (1+v) (1+m)^{a-1} -(1-v) (1-m)^{a-1} \right]. \label{dmdt} \end{equation} Equation~(\ref{dmdt}) describes the evolution of a very large system ($N \gg 1$) at the macroscopic level, neglecting finite size fluctuations. This equation for the magnetization is enough to describe the system, given that the density of neighboring nodes in opposite state $\rho$ can be indirectly obtained through the relation \begin{equation} \rho(t) = 2\, \sigma_+(t) \, \sigma_-(t) = \frac{\left[1-m^2(t)\right]}{2}. \end{equation} \subsubsection{Stability} \label{Fully-stability} \begin{figure}[t] \begin{center} \includegraphics[width=0.6\textwidth]{figure1.eps} \end{center} \caption{Coexistence and dominance regions of the Abrams-Strogatz model in a fully connected network. For values of the volatility parameter $a > 1$, the stable solutions are those of language dominance, i.e., all individuals using language $X$ ($m_s=-1$) or all using language $Y$ ($m_s=1$), whereas for $a<1$ both languages coexist, with a relative fraction of speakers that depends on $a$ and the difference between languages' prestige, measured by the bias $v$. In the extreme case $v=-1$ ($v=1$), only language switchings towards $X$ ($Y$) are allowed, and thus only one dominance state is stable, independent on $a$.} \label{stab-CG} \end{figure} Equation~(\ref{dmdt}) has three stationary solutions \begin{eqnarray} m_{-}=-1, ~~~ m^*=\frac{(1-v)^{\frac{1}{a-1}}- (1+v)^{\frac{1}{a-1}}}{(1-v)^{\frac{1}{a-1}}+(1+v)^{\frac{1}{a-1}}}~~~ \mbox{and}~~~ m_{+}=1. \label{solutions_AS} \end{eqnarray} The stability of each of the solutions depends on the values of the parameters $a$ and $v$. A simple stability analysis can be done by considering a small perturbation $\epsilon$ around a stationary solution $m_s$. For $m_s=m_{\pm}$, we replace $m$ in Eq.~(\ref{dmdt}) by $m=\pm 1 \mp \epsilon$ (with $\epsilon >0$), and expand to first order in $\epsilon$ to obtain \begin{equation} \frac{d \epsilon}{dt} = 2^{-a} \left[ (1 \mp v) \epsilon^{a-1} - 2^{a-1} (1 \pm v) \right] \epsilon. \end{equation} When $a<1$, $\epsilon^{a-1} \to \infty$ as $\epsilon \to 0$, thus both solutions $m_{\pm}$ are unstable, whereas for $a>1$, $\epsilon^{a-1} \to 0$ as $\epsilon \to 0$, thus $m_{\pm}$ are stable. In the line $a=1$, $m_{+}$ is unstable (stable) for $v<0$ ($v>0$), and vice-versa for $m_{-}$. The same analysis for the intermediate solution $m^*$ leads to \begin{eqnarray} \frac{d \epsilon}{dt} = 2^{-(a+1)} (a-1) \left( 1-{m^*}^2 \right) \left[ (1 + v) (1+m^*)^{a-2} + (1 - v) (1-m^*)^{a-2} \right] \epsilon. \nonumber \\ \label{} \end{eqnarray} Then, $m^*$ is unstable (stable) for $a>1$ ($a<1$). In Fig.~\ref{stab-CG} we show the regions of stability and instability of the stationary solutions on the $(a,v)$ plane obtained from the above analysis. We observe a region of coexistence ($m^*$ stable) and one of bistable dominance ($m_{+}$ and $m_{-}$ stable). The non-trivial stationary solution, $m^*$, is shown in Fig.~\ref{IDL_plot-stab_AS} as a function of the parameters $a$ and $v$. For the coexistence regime ($a<1$), the absolute value of the stable stationary magnetization $|m^*|$ increases with both, $|v|$ and $a$. When $v\neq0$ the coexistence solution includes a majority of agents using the language with higher prestige, and the rest of the agents using the language with lower prestige. On the contrary, for the dominance regime ($a>1$) $|m^*|$ decreases with $a$ and increases with $|v|$. \begin{figure}[t] \begin{center} \includegraphics[width=0.6\textwidth]{figure2.eps} \end{center} \caption{ Stationary solution $m^*(a,v)$ for the Abrams-Strogatz model (vertical axis) as a function of the two parameters of the model, $a$ and $v$ (horizontal-plane). See Expression~(\ref{solutions_AS}). Notice how $m^*$ approaches the values of the two trivial stationary solutions, $m_{-}=-1$ and $m_{+}=+1$ when $a\rightarrow 1$: for $v>0$, $\lim_{a \to 1^\pm} (m^*)=\mp1$. The opposite holds for $v<0$. The non-trivial stationary solution, $m^*$, is effectively not defined at $a=1$, and in this case the system has only two stationary states, $m_{-}$ and $m_{+}$. The figure illustrates the change of stability of $m^*$ at $a_c = 1$.} \label{IDL_plot-stab_AS} \end{figure} In order to account for possible finite size effects neglected in Eq.~(\ref{dmdt}) we have run numerical simulations in a fully connected network. We first notice that the solutions $m= \pm 1$ correspond to the totally ordered \emph{absorbing} configurations, that is, once the system reaches those configurations it never escapes from them. This is because, from the transition probabilities Eq.~(\ref{Ps-s}), a node never flips when it has the same state as all its neighbors. Thus, to study the stability of these solutions we have followed a standard approach \cite{Marro_Dickman_1999} that consists of adding a defect (seed) to the initial absorbing state and let the system evolve (spreading experiment). If, in average, the defect spreads over the entire system, then the absorbing state is unstable, otherwise if the defect quickly dies out, the absorbing state is stable. For instance, to study the stability of $m=-1$, we started from a configuration composed by $N-1$ down spins and $1$ up spin (seed), that corresponds to a magnetization $m=-1+2/N \gtrsim -1$, and we let the system evolve until an absorbing configuration was reached. Whether $m=-1$ is stable or not depends on the values of $v$ and $a$. If $m=-1$ is unstable, then the seed creates many up spins and spreads over the system, to end in one of the absorbing states. If $m=-1$ is stable, then the initial perturbation dies out, and the system ends in the $m=-1$ absorbing state. The theory of criticality predicts that the survival probability $P(t)$, i.e, the fraction of realizations that have not died up to time $t$, follows a power-law at the critical point \cite{Marro_Dickman_1999}, where the stability of the absorbing solution changes. Figure~\ref{P-AS} shows that for a fixed value of the bias $v=0$, $P(t)$ decreases exponentially fast to zero for values of $a>1$, while it reaches a constant value for $a<1$. For $a_{c} \simeq 1.0$, $P(t)$ decays as $P(t)\sim t^{-\delta}$, with $\delta \simeq 0.95$, indicating the transition line from an unstable to a stable solution $m=-1$ as $a$ is increased, in agreement with the previous stability analysis. \begin{figure}[t] \begin{center} \includegraphics[width=0.6\textwidth]{figure3.eps} \caption{Probability $P(t)$ that the system is still alive at time $t$, when it starts from a configuration composed by an up spin in a sea of $10^5-1$ down spins, endowed with the Abrams-Strogatz dynamics with equivalent languages (bias $v=0$), on a fully connected network. Different curves correspond to the values $a=0.90, 0.99, 1.00, 1.01, 1.10$ and $2.0$ (from top to bottom). At $a_{c} \simeq 1.0$, $P(t)$ follows a power law decay with exponent $\delta \simeq 0.95$, indicated by the dashed line.} \label{P-AS} \end{center} \end{figure} Following the same procedure, we have also run spreading experiments to check the stability transition for different values of the bias. For $v=-0.02$ and $v=-0.2$, on a system of size $N=10^5$, we found the transitions at $a \simeq 1.007$ and $a \simeq 1.052$, respectively. These values are slightly different from the analytical value $a_{c} \simeq 1.0$, but we have verified that as $N$ is increased, the values become closer to $1.0$, in agreement with the stability analysis on infinite large systems. An alternative and more visual way of studying stability in the mean field limit, is by writing Eq.~(\ref{dmdt}) in the form of a time-dependent Ginzburg-Landau equation \begin{equation} \frac{dm(t)}{dt} = - \frac{\partial V_{a,v}(m)}{\partial m}, \label{TDGL} \end{equation} with potential \begin{eqnarray} V_{a,v}(m) &\equiv& 2^{-a} \Biggl\{-v m - \frac{1}{2}(a-1) m^2 + \frac{v}{6} \left[ 2-(a-1)(a-2) \right] m^3 \nonumber \\ &+& \frac{1}{24} (a-1) \left[6-(a-2)(a-3) \right] m^4 +\frac{v}{10} (a-1)(a-2)m^5 \nonumber \\ &+& \frac{1}{36}(a-1)(a-2)(a-3) m^6 \Biggr\}. \label{Vav} \end{eqnarray} $V_{a,v}$ is obtained by Taylor expanding the term in square brackets of Eq.~(\ref{dmdt}) up to $3$-rd order in $m$, and integrating once over $m$. We assume that higher order terms in the expansion are irrelevant, and the dynamics is well described by an $m^6$-potential. Within this framework, the state of the system, represented by a point $m(t)$ in the magnetization one-dimensional space $-1<m<1$, moves ``down the potential hill'', trying to reach a local minimum. Therefore, a minimum of $V_{a,v}$ at some point $m_s$ represents a stable stationary solution, given that if the system is moved apart from $m_s$ and then released, it immediately goes back to $m_s$, whereas a maximum of $V_{a,v}$ represents an unstable stationary solution. \begin{figure}[t] \begin{center} \includegraphics[width=0.60\textwidth]{figure4.eps} \caption{Ginzburg-Landau potential from Eq.~(\ref{Vav}), for the Abrams-Strogatz model with bias $v=-0.1$ and values of volatility $a=0.8,1.0$ and $2.0$ (from top to bottom). Arrows show the direction of the system's magnetization towards the stationary solution (solid circles). For $a=0.8$ the minimum is around $m \simeq -0.5$, indicating that the system relaxes towards a partially ordered stationary state, while for $a=1.0$ and $2.0$, it reaches the complete ordered state $m=-1$.} \label{V-av} \end{center} \end{figure} As Fig.~\ref{V-av} shows, for $a<1$ and all values of $v$, the single-well potential has a minimum at $|m_s|<1$ ($m_s\simeq -0.5$ for $a=0.8$ and $v=-0.1$), indicating that the system reaches a partially ordered stable state, with fractions $0.75$ and $0.25$ of down and up spins, respectively, and a density of opposite-state links $\rho \simeq 0.375$. For $a>1$, the double-well potential has a minimum at $m=\pm 1$, thus depending on the initial magnetization, the system is driven to one of the stationary solutions $m=\pm 1$, corresponding to the totally ordered configurations in which $\rho=0$. This description works well in infinite large systems, where finite size fluctuations are zero. But in finite systems, the absorbing solutions $m=\pm 1$ are the only ``truly stationary states'', given that fluctuations ultimate take the system to one of those states. Even for the case $a<1$, where the minimum is at $|m_s|<1$, the magnetization fluctuates around $m_s$ for a very long time until after a large fluctuation it reaches $|m|=1$, and the system freezes. \subsubsection{$a=1$ case: the voter model} \label{Fully-voter} For $a=1$ the ASM becomes equivalent to the voter model. A switching probability proportional to the local density of neighbors in the opposite state is statistically equivalent to adopt the state of a randomly chosen neighbor. In this limit of neutral volatility, $a=1$, Eq.~(\ref{dmdt}) becomes \begin{equation} \frac{dm}{dt} = \frac{v}{2} (1-m^2), \end{equation} whose solution is \begin{equation} m(t) = \frac{(1+m_0) e^{v t}-(1-m_0)}{(1+m_0) e^{v t}+(1-m_0)}, \end{equation} with $m_0=m(t=0)$. For a uniform initial condition is $m_0=0$, thus \begin{equation} m(t) = \tanh(v t /2), \label{m} \end{equation} and \begin{equation} \rho(t) = \frac{1}{2} \left[ 1-\tanh^2(v t /2) \right]. \label{r} \end{equation} In Fig.~\ref{m-r-t} we observe that the analytical solutions from Eqs.~(\ref{m}) and (\ref{r}) agree very well with the results from numerical simulations of the model, for large enough systems, and they also reproduce the Monte Carlo results found in \cite{Stauffer-07}. This is so, because finite-size fluctuations effects are negligible compare to bias effects, even for a small bias. \begin{figure}[t] \begin{center} \includegraphics[width=0.60\textwidth]{figure5.eps} \caption{Abrams-Strogatz model on a fully connected network of $N=1000$ nodes with volatility $a=1$. Upper panel: Average magnetization $m$ vs time for values of the bias $v=0.8, 0.4, 0.2, 0.0, -0.2, -0.4$ and $-0.8$ (from top to bottom). Lower panel: Average density of opposite-state links $\rho$ vs time for $v=0.0, 0.2, 0.4$ and $0.8$ (top to bottom). Open symbols are the results from numerical simulations, while solid lines in the upper and lower panels correspond to the solutions from Eqs.~(\ref{m}) and (\ref{r}) respectively. Averages are over $100$ independent realizations starting from a configuration with a uniform distribution of spins and global magnetization $m(0)=0$.} \label{m-r-t} \end{center} \end{figure} When the bias is exactly zero, one obtains that in an infinite large network $dm/dt = 0$, thus $m$ and $\rho$ are conserved. However, in a finite network, fluctuations always lead the system to one of the absorbing states \cite{Stauffer-07}. To find how the system relaxes to the final state, one needs to calculate the evolution of the second moment $\langle m^2 \rangle$ of the magnetization, related to the fluctuations in $m$, where the symbol $\langle ~ \rangle$ represents an average over many realizations. This leads to a decay of the average density of opposite-state links of the form (see \cite{Vazquez-08a}) \begin{equation} \langle \rho(t) \rangle = \frac{1}{2}\left[1-\langle m^2(t) \rangle \right] = \langle \rho(0) \rangle~ e^{-2 t/N}. \end{equation} In terms of the potential description of Eq.~(\ref{TDGL}), we observe that when $v \neq 0$, $V_{a,v}$ has only one minimum (see Fig.~\ref{V-av}), thus the system has a preference for one of the absorbing states only, whereas if $v =0$, is $V_{a,v}=0$, and the magnetization is conserved ($m(t)=m(0)=$ constant). In finite systems, even though the average magnetization over many realizations is conserved, the system stills orders in individual realizations by finite size fluctuations. \subsection{Complex networks} \label{AS-nets} In real life, most individuals in a large society interact only with a small number of acquaintances, and they all form a social network of connections, where nodes represent individuals and links between them represent interactions. Thereby, we consider a network of $N$ nodes, with a given degree distribution $P_k$, representing the fraction of individuals connected to $k$ neighbors, such that $\sum_k P_k=1$. In order to develop a mathematical approach that is analytically tractable, we assume that the network has no degree correlations, as it happens for instance in Erd\"os-Renyi \cite{ER} and Barab\'asi-Albert scale-free networks \cite{Barabasi_1999}. It turns out that dynamical correlations between the states of second nearest-neighbors are negligible in voter models on uncorrelated networks \cite{Vazquez-08a,Castellano-09-b}. Thus, taking into account only correlations between first nearest-neighbors allows to use an approach, called \emph{pair approximation}, that leads to analytical results in good agreement with simulations. In this section, we shall use this approximation to build equations for the magnetization and the density of links in opposite state. In a time step $\delta t = 1/N$, a node $i$ with degree $k$ and state $s$ is chosen with probability $P_k \, \sigma_s$. Here we assume that the density of nodes in state $s$ within the subgroup of nodes with degree $k$ is independent on $k$ and equal to the global density $\sigma_s$. Then, according to transitions (\ref{PXY}), $i$ switches its sate with probability \begin{equation} P(s \to -s) = \frac{1}{2}(1-s v)\left( n_{-s}/k \right)^a, \label{Ps-s1} \end{equation} where we denote by $n_{-s}$ the number of neighbors of $i$ in the opposite state $-s$ ($0 \le n_{-s} \le k$) and the bias $v$ is defined as in the previous section. If the switch occurs, the density $\sigma_s$ is reduced by $1/N$, for which the magnetization $m=\sigma_+ - \sigma_-$ changes by $-2 s/N$, while the density $\rho$ changes by $2(k-2 n_{-s})/\mu N$, where $\mu \equiv \sum_k k P_k$ is the average degree of the network. Thus, in analogy to section \ref{Fully}, but now plugging the transition probabilities from Eq.~(\ref{Ps-s1}) into Eq.~(\ref{dmdt0}), we write the average change in the magnetization as \begin{eqnarray} \frac{d m(t)}{d t} &=& \sum_k \frac{P_k \sigma_-}{1/N} \sum_{n_{+}=0}^k B(n_{+},k) \frac{(1+v)}{2} \left( \frac{n_{+}}{k} \right)^a \frac{2}{N} \nonumber \\ &-& \sum_k \frac{P_k \sigma_+}{1/N} \sum_{n_{-}=0}^k B(n_{-},k) \frac{(1-v)}{2} \left( \frac{n_{-}}{k} \right)^a \frac{2}{N}, \label{dmdt1} \end{eqnarray} and similarly, the change in the density of links in opposite state as \begin{eqnarray} \frac{d \rho(t)}{dt} &=& \sum_k \frac{P_k \sigma_-}{1/N} \sum_{n_{+}=0}^k B(n_{+},k) \frac{(1+v)}{2} \left( \frac{n_{+}}{k} \right)^a~ \frac{2 (k-2 n_{+})}{\mu N} \nonumber \\ &+& \sum_k \frac{P_k \sigma_+}{1/N} \sum_{n_{-}=0}^k B(n_{-},k) \frac{(1-v)}{2} \left( \frac{n_{-}}{k} \right)^a~ \frac{2 (k-2 n_{-})}{\mu N}. \label{drdt1} \end{eqnarray} We denote by $B(n_{s},k)$, the probability that a node of degree $k$ and state $-s$ has $n_{s}$ neighbors in the opposite state $s$. Defining the $a$-th moment of $B(n_{s},k)$ as \begin{eqnarray*} \langle n_{s}^a \rangle_k \equiv \sum_{n_{s}=0}^k B(n_{s},k) n_{s}^a, \end{eqnarray*} we arrive to the equations \begin{equation} \frac{d m(t)}{d t} = \sum_k \frac{P_k} {2\,k^a} \left[ (1+v) (1-m) \langle n_+^a \rangle_k - (1-v) (1+m) \langle n_-^a \rangle_k \right], \label{dmdt2} \end{equation} \begin{eqnarray} \frac{d \rho(t)}{dt} &=& \sum_k \frac{P_k}{2 \, \mu \, k^{a}} \Big\{ (1+v) (1-m) \left[ k \langle n_+^a \rangle_k - 2 \langle n_+^{(1+a)} \rangle_k \right] \nonumber \\ &+& (1-v) (1+m) \left[ k \langle n_-^a \rangle_k - 2 \langle n_-^{(1+a)} \rangle_k \right] \Big\}. \label{drdt2} \end{eqnarray} \subsubsection{$a=1$ case: the voter model} In order to develop an intuition about the temporal behavior of $m$ and $\rho$ from Eqs.~(\ref{dmdt2}) and (\ref{drdt2}), we first analyze the simplest and non-trivial case $a=1$, that corresponds to the voter model on complex networks. A rather complete analysis of the time evolution and consensus times of this model on uncorrelated networks, for the symmetric case $v=0$, can be found in \cite{Vazquez-08a}. Following a similar approach, here we study the general situation in which the bias $v$ takes any value. To obtain closed expressions for $m$ and $\rho$, we consider that the system is ``well mixed'', in the sense that the different types of links are uniformly distributed over the network. Therefore, we assume that the probability that a link picked at random is of type $+ -$ is equal to the global density of $+ -$ links $\rho$. Then, $B(n_{-s},k)$ becomes the binomial distribution with \begin{equation} P(-s|s)=\rho/2 \sigma_s \label{P-ss} \end{equation} as the single event probability that a first nearest-neighbor of a node with state $s$ has state $-s$. Here, we use the fact that in uncorrelated networks dynamical correlations between the states of second nearest-neighbors vanish (pair approximation). $P(-s|s)$ is calculated as the ratio between the total number of links $\rho \mu N/2$ from nodes in state $s$ to nodes in state $-s$, and the total number of links $N \sigma_s \mu$ coming out from nodes in state $s$. Taking $a=1$ in Eqs.~(\ref{dmdt2}) and (\ref{drdt2}), and replacing the first and second moments of $B(n_{-s},k)$ by \begin{eqnarray*} \langle n_{-s} \rangle &=& P(-s|s) k, \\ \langle n_{-s}^2 \rangle &=& P(-s|s) k + P(-s|s)^2 k(k-1), \end{eqnarray*} leads to the following two coupled closed equations for $m$ and $\rho$ \begin{eqnarray} \label{dmdt3} \frac{d m(t)}{d t}&=&v \rho \\ \label{drdt3} \frac{d \rho(t)}{d t}&=&\frac{\rho}{\mu} \Bigg\{ \mu-2 - \frac{2 (\mu-1) (1+v\,m) \rho}{(1-m^2)} \Bigg\}. \end{eqnarray} For $v=0$, the above expressions agree with the ones of the symmetric voter model \cite{Vazquez-08a}. For the asymmetric case $v \ne 0$, we have checked numerically that the only stationary solutions are $(m=1, \rho=0)$ for $v>0$ and $(m=-1, \rho=0)$ for $v<0$, that correspond to the fully ordered state, as we were expecting. Even though an exact analytical solution of Eqs.~(\ref{dmdt3}) and (\ref{drdt3}) is hard to obtain, we can still find a solution in the long time limit, assuming that $\rho$ decays to zero as \begin{equation} \rho = A\, e^{-t/2\tau(v)},~~~\mbox{for}~~t \gg 1, \label{r6} \end{equation} where $A$ is a constant given by the initial state and $\tau(v)$ is another constant that depends on $v$, and quantifies the rate of decay towards the solutions $m=1$ or $m=-1$. To calculate the value of $\tau$ we first replace the ansatz from Eq.~(\ref{r6}) into Eq.~(\ref{dmdt3}), and solve for $m$ with the boundary conditions $m(\rho=0) = 1$ and $-1$, for $v>0$ and $v<0$, respectively. We obtain \begin{eqnarray} m = \left\{ \begin{array}{ll} 1-2 v \tau \rho& \mbox{if $v > 0$};\\ -1- 2 v \tau \rho & \mbox{if $v < 0$}.\end{array} \right. \end{eqnarray} Then, to first order in $\rho$ is \begin{eqnarray} (1-m^2) = \left\{ \begin{array}{ll} 4 v \tau \rho& \mbox{if $v > 0$};\\ - 4 v \tau \rho & \mbox{if $v < 0$}.\end{array} \right. \label{1m2} \end{eqnarray} Replacing the above expressions for $m$ and $(1-m^2)$ into Eq.~(\ref{drdt3}), and keeping only the leading order terms, we arrive to the following expression for $\tau$ \begin{eqnarray} \tau(v) = \left\{ \begin{array}{ll} \frac{\mu-1-v}{2 v (\mu-2)} & \mbox{if $v > 0$};\\ \frac{1- \mu - v}{2 v (\mu-2)} & \mbox{if $v < 0$}.\end{array} \right. \label{tau} \end{eqnarray} Finally, the magnetization for long times behave as \begin{eqnarray} m = \left\{ \begin{array}{ll} 1-\frac{(\mu - 1-v) A}{\mu-2} \exp{\left[-\frac{v (\mu-2)}{\mu - 1 - v}t\right]}& \mbox{if $v > 0$};\\ -1 + \frac{(\mu- 1 + v) A}{\mu-2} \exp{\left[\frac{v (\mu-2)}{\mu- 1 + v}t\right]} & \mbox{if $v < 0$}, \end{array} \right. \label{m6} \end{eqnarray} whereas the density of opposite state links decays as \begin{equation} \rho = A\, \exp{\left[-\frac{|v| (\mu-2)}{\mu- 1 - |v|}t \right]}. \label{r7} \end{equation} Using the expression for $\tau(S)$ from Eq.~(\ref{tau}) in Eq.~(\ref{1m2}), and taking the limit $v \to 0$, we find that $\rho(t) = \frac{(\mu-2)}{2(\mu-1)}\left[ 1-m(t)^2 \right]$, in agreement with previous results of the voter model on uncorrelated networks \cite{Vazquez-08a}. By taking $\mu = N-1 \gg 1$ in Eqs.~(\ref{m6}) and (\ref{r7}), we recover the expressions for $m$ and $\rho$ on fully connected networks [Eqs.~(\ref{m}) and (\ref{r}), respectively], in the long time limit. This result means that the evolution of $m$ and $\rho$ in the biased voter model on uncorrelated networks is very similar to the mean-field case, with the time rescaled by the constant $\tau$ that depends on the topology of the network, expressed by the mean connectivity $\mu$. From the above equations we observe that the system reaches the dominance state $\rho=0$ in a time of order $\tau$. For the special case $v=0$, $\tau$ diverges, thus Eqs.~(\ref{m6}) and (\ref{r7}) predict that both $m$ and $\rho$ stay constant over time. However, as mentioned in section \ref{Fully-voter}, finite-size fluctuations drive the system to the absorbing state ($\rho=0, |m|=1$). Taking fluctuations into account, one finds that the approach to the final state is described by the decay of the average density $\rho$ \cite{Vazquez-08a} \begin{equation} \langle \rho(t) \rangle = \frac{(\mu-2)}{2(\mu-1)} e^{-2 t/T}, \end{equation} where $T \equiv \frac{(\mu-1) \mu^2 N}{(\mu-1) \mu_2}$, depends on the system size $N$, and the first and second moments, $\mu$ and $\mu_2$ respectively, of the network. \subsubsection{Stability analysis} \label{diagram_AS-in_RDR} As in fully connected networks, we assume that Eq.~(\ref{dmdt2}) for the magnetization has three stationary solutions. Indeed, we have numerically verified that for different types of networks there is, apart from the trivial solutions $m=1,-1$, an extra non-trivial stationary solution $m=m^*$. Due to the rather complicated form of Eq.~(\ref{dmdt2}), we try to study the stability of the solutions in an approximate way, and find a qualitative picture of the stability diagram in the $(a,v)$ plane. For the general case in which $a$ and $v$ take any values, we assume, as in the voter model case, that $B(n_{-s},k)$ is a binomial probability distribution with single event probabilities given by Eq.~(\ref{P-ss}). Then, the explicit form for the $a$-th moment of $B(n,k)$ is \begin{equation} \langle n_{s}^a \rangle = \sum_{n_{s}=0}^k n^a C_{n_{s}}^k \left(\frac{\rho}{2 \sigma_{-s}} \right)^{n_{s}} \left( 1-\frac{\rho}{2 \sigma_{-s}} \right)^{k-n_{s}}. \label{moments} \end{equation} We also assume that, as it happens for the voter model case $a=1$ [see Eq.~(\ref{1m2})], $\rho$ and $m$ are related by $\rho(t) \simeq \frac{q}{2} \left[1-m^2(t)\right]$, where $q$ is a constant that depends on $a$ and $v$. We note that this relation satisfies the fully-ordered-state condition $\rho=0$ when $|m|=1$. We shall see that the exact functional form of $q=q(a,v)$ is irrelevant for the stability results, as long as $q>0$. To simplify calculations even more, we consider that the network is a degree-regular random graph with degree distribution $P_k = \delta_{k,\mu}$, that is, all nodes have exactly $\mu$ neighbors chosen at random. Then, replacing the above expression for the moments into Eq.~(\ref{dmdt2}), and substituting $\rho$ by the approximate value $\frac{q}{2} \left[1-m^2 \right]$, we arrive to the following closed equation for $m$ \begin{eqnarray} \frac{dm}{dt} &=& \frac{(1-m^2)}{2 \mu^a} \sum_{n=0}^\mu C_{n}^\mu\, n^a\, \left( \frac{q}{2} \right)^n \Big\{ (1+v) (1+m)^{n-1} \left[1-p(1+m)\right]^{\mu-n} \nonumber \\ &-&(1-v) (1-m)^{n-1} \left[1-p (1-m) \right]^{\mu-n} \Big\}, \label{dmdt4} \end{eqnarray} where mute indices $n_-$ and $n_+$ were replaced by the index $n$. To check the stability of $m=1$, we take $m=1-\epsilon$ in Eq.~(\ref{dmdt4}), and expand it to first order in $\epsilon$. We obtain after some algebra \begin{equation} \frac{d \epsilon}{dt} = \frac{\mu^{-a}}{2} \left( \langle n \rangle_{q} + \langle n^a \rangle_{q} \right) \left[ \mathcal V_1(a) - v \right] \epsilon, \end{equation} where the symbols $\langle ~ \rangle_{q}$ represent the moments of a Binomial distribution with probability $q$, and the bias function $\mathcal V_1(a)$ is defined as \begin{equation} \mathcal V_1(a) = \frac{\langle n \rangle_q - \langle n^a \rangle_q} {\langle n \rangle_q + \langle n^a \rangle_q}. \end{equation} Then, for a fixed value of $a$ the solution $m=1$ is stable (unstable), when $v$ is larger (smaller) than $\mathcal V_1(a)$. The shape of the function $\mathcal V_1(a)$ can be guessed using that for $a$ larger (smaller) than $1$, the moment $\langle n^a \rangle$ is larger (smaller) than $\langle n \rangle$. Then $\mathcal V_1(a)$ goes to $(\langle n \rangle -1)/(\langle n \rangle +1) \lesssim 1$ and $-1$ as $a$ approaches to $0$ and $\infty$, respectively. Also $\mathcal V_1(a)=0$, for $a=1$. With a similar stability analysis we obtained that $m=-1$ is stable (unstable) for the points $(a,v)$ below (above) the transition line $\mathcal V_{-1}(a)=- \mathcal V_1(a)$, while $m=m^*$ is stable in the region where both $m=-1$ and $m=1$ are unstable. In Fig.~\ref{stab-DR} we show a picture that summarizes the stability regions defined by the transition lines $\mathcal V_1(a)$ and $\mathcal V_{-1}(a)$. These lines were obtained by integrating numerically the two coupled Eqs.~(\ref{dmdt2}) and (\ref{drdt2}), with the moments defined in Eq.~(\ref{moments}), and finding the points $(a,v)$ where the stationary solutions $m=1,-1$ became unstable. We considered two degree-regular random graphs with degrees $\mu=3$ [solid lines $\mathcal V_1^3(a)$ and $\mathcal V_{-1}^3(a)$] and $\mu=10$ [dashed-lines $\mathcal V_1^{10}(a)$ and $\mathcal V_{-1}^{10}(a)$], thus we took $P_k = \delta_{k,\mu}$ in the equations. For clarity, only the stable solutions are labeled in the picture. We observe that as the degree of the network increases, the coexistence region expands and approaches to the corresponding region $a<1$ on fully connected networks. \begin{figure}[t] \begin{center} \includegraphics[width=0.60\textwidth]{figure6.eps} \caption{Stability diagram for the Abrams-Strogatz model on a degree-regular random graph, obtained by numerical integration of Eqs.~(\ref{dmdt2}), (\ref{drdt2}) and (\ref{moments}). The solution $m=1$ is stable above the line $\mathcal V_1$, while the solution $m=-1$ is stable below the line $\mathcal V_{-1}$. Solid and dashed lines correspond to graphs with degrees $\mu=3$ and $\mu=10$ respectively. In the coexistence region, where the stable solution is $m^*$, the system is composed by both type of users, while in the dominance region, users of either one or the other language prevail, depending on the initial state. We observe that the region of coexistence is reduced, compared to the model on fully connected networks (Fig.~\ref{stab-CG}), and that there are also two single-dominance regions where always the same language dominates.} \label{stab-DR} \end{center} \end{figure} In order to give numerical evidence, from Monte Carlo simulations, of the different phases and transition lines predicted in Fig.~\ref{stab-DR}, we have run spreading experiments as explained in section \ref{Fully}, for a degree-regular random graph (DRRG) with degree $\mu=3$ and $N=10^5$ nodes, and tested the stability of the homogeneous solutions $m=\pm 1$. We first set the bias in $v=0$ and, by varying $a$, we obtained a transition at $a_{c} \simeq 1.0$ from dominance to coexistence, as $a$ is decreased: in the dominance region the survival probability $P(t)$ decays exponentially fast to zero, indicating that $m=1$ is stable, while in the coexistence region $P(t)$ reaches a constant value larger than zero, showing that $m=1$ is unstable (not shown). This transition is the same as the one in fully connected networks (FCN) (Fig.~\ref{P-AS}). We then repeated the experiment with $v=-0.2$, whose results are summarized in Fig.~\ref{P-AS-FCN-DRRG}, where we show $P(t)$ for different values of $a$. Increasing $a$ from $0$, which corresponds to the coexistence regime ($m=\pm1$ are unstable solutions, and $m^*$ is stable), we show in Fig.~\ref{P-AS-FCN-DRRG}(a) how in a DRRG $m=-1$ changes from unstable to stable at a value $0.25 < a < 0.3$, as $P(t)$ starts to decay to zero. This corresponds to crossing the line $\mathcal V_{-1}^{3}$ in the horizontal direction (see Fig.~\ref{stab-DR}), and entering the monostable region where there exist only two solutions, $m=-1$ stable, and $m=+1$ unstable ($m^*$ becomes equal to $-1$ along the transition line $\mathcal V_{-1}^{3}$). In Fig.~\ref{P-AS-FCN-DRRG}(b) we observe how in a DRRG $m=+1$ becomes stable at a value $1.80 < a < 1.85$. This corresponds to crossing the line $\mathcal V_{1}^{3}$ (see Fig.~\ref{stab-DR}) and entering to the dominance region, where both $m=\pm 1$ are stable. Notice that for $a=1.80$, $P(t)$ first curves up and then it quickly decays to zero at a time $t \simeq 4000$. This means that a finite fraction of realizations starting from a system with a single down spin took, in average, a mean time $t \simeq 4000$ to end up in a configuration with all down spins, showing that $m=-1$ is a stable solution. This supports our claim that in the monostable region there exist only two solutions, $m=-1$ stable, and $m=+1$ unstable. These results confirm the existence of a quite broad single-dominance region in DRRG ($0.30 \lesssim a \lesssim 1.85$ for $v=-0.2$ and $\mu=3$), in agreement with the stability diagram obtained in Fig.~(\ref{stab-DR}), while this region seems to be absent in FCN. Indeed, Fig.~\ref{P-AS-FCN-DRRG}(c) shows how this unstable-stable transition happens in a FCN at a value $0.93 < a < 1.07$, in agreement with the transition line $a_{c} \simeq 1.0$ in FCN. Here, both $m=\pm 1$ gain stability at the same point, and the system enters to the dominance region (see Fig~\ref{stab-CG}). \begin{figure} \begin{center} \includegraphics[width=0.6\textwidth]{figure7.eps} \caption{Spreading experiments: probability $P(t)$ that the system is still alive at time $t$ in the Abrams-Strogatz model with bias $v=-0,2$ and various values of volatility $a$, showing the stability of the solutions $m=1,-1$. Dashed curves decay quickly to zero, indicating that the solution is stable while solid curves represent unstable solutions. (a) Degree-regular random graph (DRRG). Stability of the solution $m=-1$: $a=0.25$ (solid curve), $a=0.30$ (dashed curve). (b) Degree-regular random graph (DRRG). Stability of the solution $m=+1$: $a=1.80$ (solid curve), $a=1.85$ (dashed curve). (c) Fully connected network (FCN). Stability of the solutions $m=\pm 1$: $a=0.93$ (solid curve), $a=1.07$ (dashed curve). All curves correspond to an average over $10^5$ independent realizations on networks with $N=10^5$ nodes.} \label{P-AS-FCN-DRRG} \end{center} \end{figure} In summary, we find that, compared to the fully connected case, the region of coexistence is shrunk for $v\neq0$, as there appear two regions where only one solution is stable. These regions also reduce part of the dominance region. The effect of the bias is shown to be more important in DRRGs with low connectivity $\mu$ and, as a general result, coexistence becomes harder to achieve in sparse networks. \section{Bilinguals Model} \label{MW-model} This model can be regarded as an extension of the Abrams-Strogatz model in which, besides monolingual users $X$ and $Y$, there is a third class of individuals that use both languages, that is, bilingual users labeled with state $Z$. A monolingual $X$ ($Y$) becomes a bilingual with a rate depending on the number of its neighbors that are monolinguals $Y$ ($X$), while direct transitions from one class of monolingual to the other are forbidden. This reflects the fact that individuals that use one language only, are forced to start using both languages if they want to have a conversation with monolingual users of the opposite language. For a similar reason, the transition from a bilingual $Z$ to a monolingual $X$ ($Y$) depends on the number of neighbors using language $X$ ($Y$), which includes bilingual agents. Thus, the transition probabilities between states are given by \begin{eqnarray} P(X \to Z) &=& (1-\mathcal S) \, \sigma_y^a, \nonumber \\ P(Z \to Y) &=& (1- \mathcal S) \, (1-\sigma_x)^a, \nonumber \\ P(Y \to Z) &=& \mathcal S \, \sigma_x^a, \nonumber \\ P(Z \to X) &=& \mathcal S \, (1-\sigma_y)^a, \label{PXYZ} \end{eqnarray} where $\sigma_x$, $\sigma_y$ and $\sigma_z$ are the densities of neighboring speakers in states $X$, $Y$ and $Z$ respectively, and $\mathcal S$ is the prestige of language $X$. As in the ASM, it is convenient to consider monolinguals $X$ and $Y$, as particles with opposite spins $-1$ and $1$ respectively. Bilinguals are considered as spin-$0$ particles because they are a combination of the two opposite states. Given that the model is invariant under the interchange of $-1$ and $1$ particles, the system is better described using the global magnetization $m \equiv \sigma_+ - \sigma_-$ and the density of bilinguals $\sigma_0$, where $\sigma_-$,$\sigma_0$, $\sigma_+$, are the global densities of nodes in states $-1$, $0$ and $1$, respectively. Another alternative could be the use of the density of connections between different states $\rho \equiv 2 \sigma_- \sigma_+ + 2 \sigma_- \sigma_0 + 2 \sigma_+ \sigma_0$, but numerical simulations show that $\rho$ and $\sigma_0$ are proportional. We now study the evolution of the system on fully connected and complex networks, by writing equations for $m$ and $\sigma_0$. \subsection{Fully connected networks} In the fully connected case, the local densities of neighbors in the different states agree with the global densities $\sigma_-$,$\sigma_0$, $\sigma_+$, thus, using the transition probabilities Eqs.~(\ref{PXYZ}), the rate equations for $\sigma_-$ and $\sigma_+$ can be written as \begin{eqnarray} \frac{d \sigma_-}{dt} &=& \frac{(1-v)}{2} \sigma_0 (1-\sigma_+)^a - \frac{(1+v)}{2} \sigma_- \sigma_+^a, \\ \frac{d \sigma_+}{dt} &=& \frac{(1+v)}{2} \sigma_0 (1-\sigma_-)^a - \frac{(1-v)}{2} \sigma_+ \sigma_-^a, \end{eqnarray} where $v \equiv 1-2 \mathcal S$ is the bias. The rate equations for $m=\sigma_+-\sigma_-$ and $\sigma_0=1-\sigma_+-\sigma_-$ can be derived from the above two equations, and by making the substitutions $\sigma_s = (1-\sigma_0 + s\, m)/2$, with $s=\pm 1$. We obtain \begin{eqnarray} \label{dmdt5} \frac{dm}{dt} &=& 2^{-(2+a)} \Big\{ 2 \sigma_0 \left[ (1+v) (1+\sigma_0+m)^a - (1-v) (1+\sigma_0 -m)^a \right] \\ &+& (1+v) (1-\sigma_0-m) (1-\sigma_0 +m)^a - (1-v) (1-\sigma_0+m) (1-\sigma_0-m)^a \Big\} \nonumber \end{eqnarray} and \begin{eqnarray} \label{dsdt} \frac{d\sigma_0}{dt} &=& 2^{-(2+a)} \Big\{ -2 \sigma_0 \left[ (1+v) (1+\sigma_0+m)^a +(1-v) (1+\sigma_0 -m)^a \right] \\ &+& (1+v) (1-\sigma_0-m) (1-\sigma_0 +m)^a + (1-v) (1-\sigma_0+m) (1-\sigma_0-m)^a \Big\}. \nonumber \end{eqnarray} Equations~(\ref{dmdt5}) and (\ref{dsdt}) are difficult to integrate analytically, but an insight on its qualitatively behavior can be obtained by studying the stability of the stationary solutions with $a$ and $v$. As in the ASM, we expect that, for a given $v$, an order-disorder transition appears at some value $a_c$ of the volatility parameter, where the stability of the stationary solutions changes. If $a$ is small, then flipping rates are high, thus we expect the system to remain in an active disordered state, while for large enough values of $a$ spins tend to be aligned, thus the system should ultimately reach full order. We now calculate the transition point for the symmetric case $v=0$, and then find an approximate solution for the linear case $a=1$. \begin{figure}[t] \begin{center} \includegraphics[width=0.5\textwidth]{figure8.eps} \caption{Spreading experiments: probability $P(t)$ that the system is still alive at time $t$ in the Bilinguals model on a fully connected network, obtained from the same spreading experiments and parameters ($v=0$, $N=10^5$) as described in Fig.~\ref{P-AS} for the Abrams-Strogatz model. The curves correspond to volatilities $a=0.600, 0.618, 0.620, 0.622$ and $0.700,$ (top to bottom). $P(t)$ decays as $t^{-\delta}$ at the transition point $0.620$ (close to the theoretical value $a_{fc} \simeq 0.63$), with $\delta \simeq 1.76$, indicated by the dashed line.} \label{P-MW} \end{center} \end{figure} \subsubsection{Transition point for $v=0$} In the symmetric case $v=0$ one can easily verify that the points $(m=\pm 1, \sigma_0=0)$ in the $(m,\sigma_0)$ plane are two stationary solutions of Eqs.~(\ref{dmdt5}) and (\ref{dsdt}). But there is also a third non-trivial stationary solution $(m=0, \sigma_0 = \sigma_0^*)$, where $\sigma_0^*$ satisfies \begin{equation} 2 \sigma_0^* (1+\sigma_0^*)^a - (1-\sigma_0^*)^{(1+a)}=0. \label{stat} \end{equation} By doing a small perturbation around $(0,\sigma_0^*)$ in the $\sigma_0$ direction, one finds from Eq.~(\ref{dsdt}) that the point $(0,\sigma_0^*)$ is stable for all values of $a$. Instead, the stability in the $m$ direction changes at some value $a_{fc}$ ($fc$ stands for fully connected). Replacing $m$ by $\epsilon \ll 1 $ and $\sigma_0$ by $\sigma_0^*$ in Eq.~(\ref{dmdt5}), one arrives to the following relation that $\sigma_0^*$ and $a_{fc}$ hold when the stability changes \begin{equation} 2\, a_{fc}\, \sigma_0^* (1+\sigma_0^*)^{(a_{fc}-1)} + (a_{fc}-1) (1-\sigma_0^*)^{a_{fc}} = 0. \label{stab} \end{equation} Combining Eqs.~(\ref{stat}) and (\ref{stab}), one arrives to the following closed equation for $a_{fc}$ \begin{equation} a_{fc}~\ln { \left( \frac{1-a_{fc}}{a_{fc}} \right) } = \ln { \left( \frac{2a_{fc}-1}{1-a_{fc}} \right)}, \end{equation} whose solution is $a_{fc} \simeq 0.63$. Then, assuming that the transition point does not depend on $v$ for FCN, as it happens in the ASM, we find that the $(a,v)$ plane is divided into two regions. In the region $a<a_{fc}$, the stable solution is $(0,\sigma_0^*)$, representing a stable mix of the three kinds of individuals, while in the region $a>a_{fc}$, the stable solutions $(\pm 1,0)$ indicate the ultimate dominance of one of the languages. By performing spreading experiments we estimated that the transition point for a network of $N=10^5$ nodes is around $a = 0.62$ (see Fig.~\ref{P-MW}), and we observed that this value approaches to the analytical one $a_{fc} \simeq 0.63$ as $N$ increases. We have also checked numerically the transition point when there is a bias ($v \neq 0$), that is, when the two languages are not equivalent. In this case we found a transition around $a = 0.675$ for a bias $v=-0.2$ and $N=10^5$ nodes, what represents a small deviation from $a_{fc}$. However this difference is similar to the one found for the ASM in Section~\ref{Fully-stability} with the same system size. Therefore, we assume that this discrepancy is again due to finite size effects and, in the thermodynamic limit, the transition should be at $a_{fc}$, for any value of $v$. We note that the transition point $a_{fc} \simeq 0.63$ is smaller than the corresponding value $a_c \simeq 1.0$ for the ASM, thus the region for coexistence is reduced in the BM. This has a striking consequence. Suppose that there is population with individuals that can use only one of two languages at a time, and it is characterized by a volatility $a=0.8$, that allows the stable coexistence of the two languages. If now the behavior of the individuals is changed, so that they can use both languages before they start using the opposite language, the population looses the coexistence and finally approaches to a state with the complete dominance of one language. In other words, within these models, bilinguals in use hinder language coexistence. \subsubsection{AB Model: Neutral volatility and symmetric case} For $a=1$ and $v=0$, Eqs.~(\ref{dmdt5}) and (\ref{dsdt}) are reduced to \begin{equation} \frac{dm}{dt} = \frac{1}{2} \sigma_0 m, \label{dmdt6} \end{equation} \begin{equation} \frac{d \sigma_0}{dt} = \frac{1}{4} (1-m^2-4 \sigma_0 - \sigma_0^2). \label{dsdt1} \end{equation} The three stationary solutions are $(m,\sigma_0) = (-1, 0)$; $(1,0)$ and $(0,\sqrt{5}-2)$. Given that the above equations are difficult to integrate analytically, we try an a approximate solution by assuming that the density of bilinguals is proportional to the interface density $\rho$, something observed in our simulations, and already found in \cite{Castello-06} for the \emph{AB-model} (equivalent to the BM in the case $v=0$ and $a=1$). Bilinguals are at the interface between monolinguals, for all the networks studied. Then we write $\sigma_0 \simeq \alpha \rho$, where $\alpha$ is a constant and $\rho = 2 \sigma_- \sigma_+ + 2 \sigma_0 (\sigma_-+\sigma_+) = \frac{1}{2} \left[ (1-\sigma_0)^2 - m^2 \right] + 2 \sigma_0 (1-\sigma_0)$, from where we obtain that $m$ can be expressed in terms of $\sigma_0$ as $m^2=(1-\sigma_0)^2+4 \sigma_0 (1-\sigma_0) - 2 \sigma_0/\alpha$. Replacing this expression for $m^2$ into Eq.~(\ref{dsdt1}), we obtain the following equation for $\sigma_0$ \begin{equation} \frac{d \sigma_0}{dt} = \frac{\sigma_0}{2} (-3+\frac{1}{\alpha} + \sigma_0). \end{equation} We have checked by numerical simulations that $\alpha > 1/3$, then the solution of the above equation in the long time limit is $\sigma_0 \sim e^{(-3+1/\alpha)t/2}$. Therefore, $\sigma_0$ and $|m|$ approach to $0$ and $1$, respectively, and the system reaches full order exponentially fast. \bigskip \subsection{Complex networks} We now consider the model on complex networks. Following the same approach as in Section \ref{AS-nets}, it is possible to write down a set of nine coupled differential equations: three for the densities $\sigma_-$, $\sigma_0$ and $\sigma_+$ of node states, and six for the densities $\rho_{--}$, $\rho_{-0}$, $\rho_{+0}$, $\rho_{+-}$, $\rho_{+0}$ and $\rho_{++}$ of different types of links. However, due to the complexity of these equations, we have limited our study to the investigation of the stability regions through Monte Carlo simulations. We found that in a degree-regular random graph with mean degree $\mu=3$, the stability diagram is qualitatively similar to the one in Fig.~\ref{stab-DR} for the ASM, where the coexistence region corresponds to stationary states with a mix of the three types of speakers. Also, the coexistence-dominance transition point for $v=0$ is at $a_{cn} \simeq 0.3$ ($cn$ stands for complex networks). For $v=-0.02$, a monostable region appears for $0.2 \lesssim a \lesssim 0.4$, while this region becomes wider for $v=-0.2$ ($0 \lesssim a \lesssim 1.4$). We have also observed that the coexistence region disappears for $|v|\geq0.2$. Therefore, in the BM, the region for coexistence also shrinks as the connectivity of the network decreases (going from fully connected to complex networks with low degree), but on top of that, there exists a shift of the critical value from $a_{fc} \simeq 0.63$ (fully connected networks) to $a_{cn} \simeq 0.3$ (degree-regular random graphs). In summary, compared to the ASM, the overall effect of the inclusion of bilingual agents is that of a large reduction of the region of coexistence. \section{Square lattices} \label{Square} Dynamical properties of the ASM and BM in square lattices can be explored for different initial conditions, system sizes, and values of the prestige and volatility parameters, through a simulation applet available online \cite{languageApplet}. It turns out that the behavior of these models in square lattices is very different to their behavior in fully connected or complex networks. On the one hand, the mean distance between two sites in the lattice grows linearly with the length of the lattice side $L$, thus a spin only ``feels'' the spins that are in its near neighborhood, and therefore the mean-field approach that works well in fully connected networks gives poor results in lattices. On the other hand, correlations between second, third and higher order nearest-neighbors are important in lattices, what causes the formation of same-spin domains, unlike in random networks where correlations to second nearest-neighbors are already negligible. Thus, pair approximation does not provide a good enough description of the dynamics in lattices either, and one is forced to implement higher order approximations (triplets, quadruplets, etc), that lead to a coupled system of many equations, impossible to solve analytically. Due to the fact that the mean-field and pair approximations, that use global quantities such as the magnetization and the density of opposite-state links to describe the system, do not give good results in lattices, we follow here a different approach to obtain a macroscopic description. This approach, also developed in \cite{Vazquez-08c} for general nonequilibrium spin models consists in deriving a macroscopic equation for the evolution of a continuous space dependent spin field. Within this approach it is possible to describe coarsening processes, that is, processes of growth of local linguistic domains caused by the motion of linguistic boundaries (interface motion). In particular, one can explain whether the system orders or not, or if the ordering is curvature driven (interface motion due to surface tension reduction) or noise driven (without surface tension). We focus here on the ASM, but this macroscopic description can also be applied for systems with three states, as the BM (see \cite{Dall'Asta-08}). Given that neighboring spins tend to be aligned -due to the ferromagnetic nature of the interactions-, and also correlations between spins reinforce the alignment between far neighbors, the dynamics is characterized by the formation of same-spin domains. Starting from a well-mixed system with up and down spins randomly distributed over the lattice, after a small transient, if we look at the lattice from far we see domains growing and shrinking slowly with time, and we can interpret this dynamics at the coarse-grained level as the evolution of a continuous \emph{spin field} $\phi$ over space and time. Then, we define by $\phi_{\bf r}(t)$ the spin field at site ${\bf r}$ at time $t$, which is a continuous representation of the spin at that site ($-1<\phi<1$), also interpreted as the average value of the spin over many realizations of the dynamics. Thus, we assume that there are $\Omega$ spin particles at each site of the lattice, and we replace $\phi_{\bf r}(t)$ by the average spin value $\phi_{\bf r}(t) \to \frac{1}{\Omega} \sum_{j=1}^\Omega S_{\bf r}^j$, where $S_{\bf r}^j$ is the spin of the $j$-th particle inside site ${\bf r}$. Within this formulation, the dynamics is the following. In a time step of length $\delta t = 1/\Omega$, a site ${\bf r}$ and a particle from that site are chosen at random. The probability that the chosen particle has spin $s=\pm 1$ is equal to the fraction of $\pm$ spins in that site $(1 \pm \phi_{\bf r})/2$. Then the spin flips with probability \begin{equation} P(s \to -s) = \frac{1}{2} (1- s v) \left(\frac{1 - s \psi_{\bf r}} {2}\right)^a, \label{Ps-s2} \end{equation} where $\psi_{\bf r} \to \frac{1}{4} \sum_{{\bf r'/r}}\phi_{\bf r'}(t)$ is the average neighboring field of site ${\bf r}$, and the sum is over the $4$ first nearest-neighbors sites ${\bf r'}$ of site ${\bf r}$. If the flip happens, $\phi_{\bf r}$ changes by $-2 s/\Omega$, thus its average change in time is given by the rate equation \begin{equation} \frac{\partial \phi_{\bf r}(t)}{\partial t} = \left[1-\phi_{\bf r}(t)\right] P(- \to +) - \left[1+\phi_{\bf r}(t)\right] P(+ \to -), \label{dphi-dt} \end{equation} where the first (second) term corresponds to a $- \to +$ ($+ \to -$) flip event. In order to obtain a closed equation for $\phi$ (see \ref{Eq-phi} for details), we substitute the expression for the transition probabilities Eq.~(\ref{Ps-s2}) into Eq.~(\ref{dphi-dt}), we then expand around $\psi_{\bf r}=0$, and replace the neighboring field $\psi_{\bf r}$ by $\phi_{\bf r} + \Delta \phi_{\bf r}$, where $\Delta$ is defined as the standard Laplacian operator $\Delta \phi_{\bf r} \equiv \frac{1}{4} \sum_{\bf r'/r} \left( \phi_{\bf r'}-\phi_{\bf r} \right) =\psi_{\bf r}-\phi_{\bf r}$. Keeping the expansion up to first order in $\Delta \phi_{\bf r}$, results in the following equation for the spin field \begin{eqnarray} \frac{\partial \phi_{\bf r}(t)}{\partial t} &=& 2^{-a} \left(1-\phi_{\bf r}^2\right) \Biggl[ v+(a-1)\phi_{\bf r}+\frac{v}{2} (a-1)(a-2)\phi_{\bf r}^2 \nonumber \\ &+& \frac{1}{6} (a-1)(a-2)(a-3) \phi_{\bf r}^3 \Biggr] \nonumber \\ &+& 2^{-a} a \left[ 1 + v (a-2) \phi_{\bf r} + \frac{1}{2} (a-1)(a-4) \phi_{\bf r}^2 \right] \Delta \phi_{\bf r}. \label{dphi-dt2} \end{eqnarray} Equation~({\ref{dphi-dt2}) can be written in the form of a time dependent Ginzburg-Landau equation \begin{equation} \frac{\partial \phi_{\bf r}(t)}{\partial t} = D(\phi_{\bf r}) \Delta \phi_{\bf r} - \frac{\partial V_{a,v}(\phi_{\bf r})}{\partial \phi_{\bf r}}, \end{equation} with diffusion coefficient \begin{eqnarray} D(\phi_{\bf r}) \equiv 2^{-a} a \left[ 1 + v (a-2) \phi_{\bf r} + \frac{1}{2} (a-1)(a-4) \phi_{\bf r}^2 \right] \end{eqnarray} and potential \begin{eqnarray} V_{a,v}(\phi_{\bf r}) &\equiv& 2^{-a} \Biggl\{-v \phi_{\bf r} - \frac{1}{2}(a-1) \phi_{\bf r}^2 + \frac{v}{6} \left[ 2-(a-1)(a-2) \right] \phi_{\bf r}^3 \nonumber \\ &+& \frac{1}{24} (a-1) \left[6-(a-2)(a-3) \right] \phi_{\bf r}^4 +\frac{v}{10} (a-1)(a-2)\phi_{\bf r}^5 \nonumber \\ &+& \frac{1}{36}(a-1)(a-2)(a-3) \phi_{\bf r}^6 \Biggr\}, \end{eqnarray} which is analogous to the potential for the global magnetization $m$ in the fully connected network case (Fig.~\ref{V-av}). As we already discussed in section \ref{Fully}, for the asymmetric case $v \not= 0$ the ordering dynamics is strongly determined by $v$. When $a>1$, $V_{a,v}$ has the shape of a double-well potential with minima at $\phi = \pm 1$, and with a well deeper than the other, thus the system is quickly driven by the bias towards the lowest minimum, reaching full order in a rather short time. For $a<1$ there is a minimum at $|\phi|<1$, thus the system relaxes to a partially ordered state of language coexistence composed by a well mixed population with different proportions of speakers of the two languages. \begin{figure}[t] \begin{center} \includegraphics[width=0.6\textwidth]{figure9.eps} \caption{Ginzburg-Landau potential Eq.~(\ref{Va}) for the symmetric case $v=0$ of the Abrams-Strogatz model, with volatility values $a=0.5, 0.8, 1.0, 1.2$ and $2.0$ (from top to bottom). For $a=0.5$ and $0.8$ the system relaxes to an active state with the same fraction of up and down spins uniformly distributed over the space, corresponding to the minimum of the potential at $\phi=0$, while for $a=1.2$ and $2.0$ it reaches full order, described by the field $|\phi|=1$.} \label{V-a} \end{center} \end{figure} Specially interesting is the analysis of the symmetric case $v=0$, for which the potential is (see Fig.~\ref{V-a}) \begin{eqnarray} V_a(\phi_{\bf r}) = 2^{-a} (a-1) \biggl\{-\frac{\phi_{\bf r}^2}{2}+ \left[6-(a-2)(a-3)\right] \frac{\phi_{\bf r}^4}{24}+(a-2)(a-3)\frac{\phi_{\bf r}^6}{36} \biggr\}. \nonumber \\ \label{Va} \end{eqnarray} In this bias-free case, when $a<1$ the minimum is at $\phi=0$, thus the average magnetization in a small region around a given point ${\bf r}$ is zero, indicating that the system remains disordered (language coexistence). This can be seen in Fig.~\ref{lattice}(b), where we show a snapshot of the lattice for the model with $v=0$ and $a=0.5$, after it has reached a stationary configuration. For $a > 1$ the potential has two wells with minima at $\phi=\pm 1$, but with the same depth, thus there is no preference for any of the two states, and the system orders in either of the language dominance states by spontaneous symmetry breaking. The order-disorder nonequilibrium transition at $a=1$ is reminiscent of the well known Ising model transition, but with the volatility parameter $a$ playing the role of temperature: high volatility $a<1$ corresponds to the high temperature paramagnetic phase and low volatility to the low temperature phase. An important difference is that the transition is here first order, since the low volatility stable sates $\phi=\pm 1$ appear discontinuously at $a=1$. In addition, while in the low temperature phase of the Ising model, spins flip in the bulk of ordered domains by thermal fluctuations, here, spin flips in the low volatility regime only occur at the interfaces (domain boundaries). Complete ordering for $a > 1$ is achieved through domain coarsening driven by surface tension \cite{Gunton_1983}. That is, as the system evolves, same-spin domains are formed, small domains tend to shrink and disappear while large domains tend to grow. Figure~\ref{lattice}(d) shows a snapshot of the lattice for the evolution of the model with $a=2$. We observe that domains have rounded boundaries given that the dynamics tends to reduce their curvature, leading to an average domain length that grows with time as $l \sim t^{1/2}$ \cite{Castello-06,Dall'Asta-08}. For the special case $a=1$ (voter model) the potential is $V_a=0$, there is still coarsening but without surface tension, meaning that domain boundaries are driven by noise, as seen in Fig.~\ref{lattice}(c). As a consequence of this, the average length of domains grows very slowly with time, as $l \sim \ln t$ \cite{Krapivsky-92,Frachebourg-96,Redner-01}. \begin{figure}[t] \begin{center} \begin{tabular}{|c|c|} \hline \includegraphics[width=1.5in, bb=70 70 550 550]{figure10a.eps} & \includegraphics[width=1.5in, bb=70 70 550 550]{figure10b.eps} \\(a) & (b) \\ \hline \includegraphics[width=1.5in, bb=70 70 550 550]{figure10c.eps} & \includegraphics[width=1.5in, bb=70 70 550 550]{figure10d.eps} \\(c) & (d) \\ \hline \end{tabular} \caption{Snapshots of the Abrams-strogatz model with bias $v=0$ on a $128 \times 128$ square lattice and three values of volatility, $a=0.5$ (b), $a=1.0$ (c) and $a=2.0$ (d). (a) Initial state: each site is occupied with a spin $+1$ or $-1$ with the same probability $1/2$. (b) The system reaches an active disordered stationary state, with a global magnetization that fluctuates around zero. (c) The system displays coarsening driven by noise, characterized by domains with noisy boundaries. (d) There is also coarsening but driven by surface tension, generating domains with more rounded boundaries.} \label{lattice} \end{center} \end{figure} In order to compare the behavior of the language competition models described before in fully connected and complex networks with their behavior in square lattices we have numerically explored the stability regions in the $(a,v)$ plane for the ASM and BM in square lattices. The coexistence-dominance transition in the ASM for $v=0$ is at $a_c \simeq 1.0$, as in fully connected and complex networks, whereas the region for coexistence is found to be much more narrow than the ones observed in complex networks with low degree, like the one depicted in Fig.~\ref{stab-DR} for $\mu=3$. Using the simulation applet \cite{languageApplet} one can check that for a given value of $v \not=0$, the disordered stationary state that characterizes coexistence is harder to maintain in square lattices than in random networks: in order to have an equivalent situation, a smaller value of $a$ is needed in the former case. In the BM, apart from the narrowing of the coexistence region, we also found that the transition point for $v=0$ is shifted to an even smaller value of the volatility $a$ than in complex networks. To see this, in Fig.~\ref{r-MW} we show the time evolution of the inverse of the average interface density $\langle \rho \rangle$ \footnote{The $\langle...\rangle$ indicates average over independent realizations of the dynamics with different random initial conditions.} for various values of $a$, on a square lattice of size $N=400^2$. We observe that $\langle \rho \rangle$ decays to zero for values of $a>0.16$, indicating that the system orders (dominance phase), while $\langle \rho \rangle$ approaches to a constant value larger than zero for $a<0.16$, thus the system remains disordered (coexistence phase). At the transition point $a_{sl} \simeq 0.16$ we have that $\langle \rho \rangle \sim 1/\ln(t)$, indicating that the transition belongs to the Generalized Voter class, a typical transition observed in spin systems with two symmetric absorbing states \cite{Al_Hammal_Chate-05,Vazquez-08c,Dornic2001,Droz2003}. The fact that $a_{sl} \simeq 0.16 $ is smaller than the corresponding transition points $a_{fc}\simeq 0.63$ and $a_{cn} \simeq 0.3$, together with the narrowing effect mentioned above, leads to the result that the region for coexistence is largely reduced in square lattices, compared to fully connected and complex networks. \begin{figure}[t] \begin{center} \includegraphics[width=0.6\textwidth]{figure11.eps} \caption{Inverse of the average interface density $\langle \rho \rangle$ vs time, on a log-linear scale, for the Bilinguals model. From top to bottom: $a=0.30$, $0.20$, $0.17$, $0.16$, $0.15$ and $0.10$. Averages were done over $10^3$ independent realizations on a square lattice of side $L=400$. $\langle \rho \rangle$ decays as $1/\ln(t)$ at the transition point $a_{sl} \simeq 0.16$ (solid squares), corresponding to the behavior of a Generalized voter transition in two dimensions.} \label{r-MW} \end{center} \end{figure} \section{Summary and conclusions} \label{Summary} We have discussed the order-disorder transitions that occur in the volatility-prestige parameter space of two related models of language competition dynamics, the Abrams-Strogatz and its extension to account for bilingualism: the Bilinguals Model. We have analyzed their microscopic dynamics on fully connected, complex random networks and two-dimensional square lattices and constructed macroscopic descriptions of these dynamics accounting for the observed transitions. At a general level, we have found that both models share the same qualitative behavior, showing a transition from coexistence to dominance of one of the languages at a critical value of the volatility parameter $a_c$. The fact that agents are highly volatile ($a<a_c$), i.e, loosely attached to the language they are currently using, leads to the enhancement of language coexistence. On the contrary, in a low volatility regime ($a>a_c$), the final state is one of dominance/extinction. A more detailed comparison of both models shows important differences: In the mean field description for fully connected networks, and for the ASM, a scenario of coexistence is obtained for $a < 1$. This is independent of the relative prestige of the languages, $v$. However, the stationary fraction of agents in the more prestigious language increases with a higher prestige. But when bilingual agents are introduced (BM), the scenario of coexistence becomes the parameter space area corresponding to $a<0.63$. That is, the area of coexistence is reduced: agents with a higher level of volatility (smaller a) are needed in order to obtain a coexistence regime. Within this current framework, allowing the agents to use two languages at the same time, which is reasonable from a sociolinguistic point of view, has the effect of making language coexistence more difficult to achieve, in the sense that coexistence occurs for a smaller range of parameters. Network topology and local effects have been addressed through pair approximations for degree uncorrelated networks. For the ASM on degree-regular random networks we find that the decrease of the network connectivity leads to a reduction, in the parameter space ($a, v$), of the area of language coexistence and the area of bistable dominance, while monostable dominance regions appear, in which only the state of dominance of the more prestigious language is stable. To gain intuition on this result, we first notice that in the fully connected network, the area of coexistence ($a<1$) corresponds to a situation in which the majority of the agents use the more prestigious language. The fact that all agents are interconnected, translates to a situation in which users of the less prestigious language (minority) are in contact with every other agent in the network. In this situation, high volatility (agents switching their language use easily) is effective in order to achieve a steady state situation with individuals continuously changing the use of their language and making coexistence possible. In contrast, when considering a degree-regular random network, that is, when limiting the number of neighbors in a society, the existence of bias ($v\neq0$) opens the possibility for agents in the majority language to be placed in domains without contact with the minority language. For a region of the parameter space where there is coexistence in a fully connected network, these domains can grow in size in a random network until they occupy the entire system. This gives rise to the monostable region of dominance of the more prestigious language found in complex networks with low connectivity. Compared to the fully connected case, a higher volatility is needed in order to overcome this topological effect, leading to a reduction of the area of coexistence. In two dimensional square lattices the coexistence is shown to be even more difficult to achieve, probably due to the fact that correlations with second neighbors make the coarsening process of formation and growth of domains easier. The macroscopic field description introduced for square lattices accounts for the different coarsening processes observed for large and small volatility. The network effects described above for the ASM are also qualitatively valid for the BM. However, the reduction of the area of language coexistence is more important when considering bilingual agents. We find a shift of the critical value with the topology: $a_{fc} \simeq 0.63$ in fully connected networks, $a_{cn} \simeq 0.3$ in complex uncorrelated networks, and $a_{sl} \simeq 0.16$ in two dimensional square lattices. In summary, building upon previous works on language competition \cite{Abrams-03, Wang-05_TRENDS_Ecology, Minett-08}, we have studied numerically and by analytical macroscopic descriptions, two microscopic models for the dynamics of language competition. We have analyzed the role of bilingual agents and social network structure in the order-disorder transitions occurring for different values of the two parameters of the models: the relative prestige of the languages and the volatility of the agents. We have found that the scenario of coexistence of the two languages is reduced when bilingual agents are considered. This reduction also depends on the social structure, with the region of coexistence shrinking when the connectivity of the network decreases. We acknowledge support from project FISICOS (FIS2009-60327) of MEC and FEDER, and from NEST-Complexity project PATRES (043268). \bigskip
\section{INTRODUCTION} Since the influential paper of Ho et al. (1995) there has been a strong impetus to develop mathematical models for better understanding the interaction between HIV and the immune system; see Nowak and May (2000). However the statistical inference in these models has raised major challenges coming from the intrication of identifiability and numerical problems. The first problem is numerical: in general the trajectories of the interesting quantities (e.g. viral load or CD4 counts) are solutions of non-linear differential equations that do not have analytical solutions. The second problem is the identifiability problem: the observations recorded on one subject are not informative enough to estimate all the parameters of the model. The first problem is either avoided, simplifying the models to obtain analytical solutions (Wu and Ding, 1999), or solved by using numerical solvers of ordinary differential equations (ODE); Ramsay et al. (2007) proposed an original approach but did not apply it to a random effect model. The second problem is partly treated by considering that the particular values of the parameters for each subject are realizations of random variables with a given distribution in the population. This puts the problem in the framework of non-linear mixed effects models. Laplace approximation of the numerical integrals involved in the computation of the likelihood has been proposed (Beal and Sheiner, 1982; Lindstrom and Bates, 1990); adaptive Gaussian quadrature is another possibility (see Davidian and Giltinan, 1995). We refer to Wu (2005) for a review of statistical issues in HIV models. Recently a stochastic approximation EM (SAEM) algorithm has been proposed (Kuhn and Lavielle, 2005; Donnet and Samson, 2007). In the specific case of HIV dynamics models a Bayesian approach has been proposed by Putter et al. (2002) and Huang, Liu and Wu (2006), while a special algorithm for computing the likelihood and maximizing using a Newton-like method has been proposed by Guedj, Thi\'ebaut and Commenges (2007). However all these methods present difficulties and can be time-consuming. The hierarchical likelihood (h-likelihood) has been proposed for generalized linear models with random effects by Lee and Nelder (1996) and further studied in Lee and Nelder (2001) and Lee, Pawitan and Nelder (2006) and for non-linear mixed effects models by Noh and Lee (2008). This is very similar to an approach called penalized likelihood used by McGilChrist and Aisbett (1991) and Therneau and Grambsch (2000) for frailty models. The main idea is to treat the random effects (or the frailties) as parameters and to find estimates of all the parameters by maximizing a function which is essentially the loglikelihood conditional on the random effects minus a penalty term which takes large values if the ``random'' parameters are very dispersed. Penalized likelihood has also been used for function estimation (O'Sullivan; 1988). The advantage of this approach is that it may avoid computing numerical integrals. The curse of dimensionality is transferred from the dimension of numerical integrals to the dimension of the space on which maximization takes place. There are problems with this approach. One is the asymptotic distribution of the estimators of the fixed parameters; another is the estimation of the variances of the random parameters. Consistency of the maximum h-likelihood estimators (MHLE) has not been proved. It is often suggested to revert to the likelihood to have consistent estimators of the fixed parameters, but then the most important benefits of h-likelihood in terms of computational burden is lost. Last but not least is the problem of maximizing a complicated function over several hundred parameters. The aim of this paper is to develop a (partly non-standard) h-likelihood approach to HIV dynamics models which completely avoids computation of the likelihood. This is in the spirit of penalized likelihood in the sense that we do not try to precisely estimate the variances of the random effects. One aim is to study the asymptotic distribution of the MHLE for a given choice of the penalty. Another aim is to find an efficient maximization algorithm. The paper is organized as follows. In section 2 we describe a statistical model based on an ODE system in a general form and in a particular form which will be used for simulations. In section 3 we describe h-likelihood and we give the asymptotic distribution of the MHLE for fixed effects when the number of subjects tends toward infinity. We propose a parametric bootstrap procedure to correct the bias of the MHLE. In section 4 we propose a strategy for choosing the penalty based on the guess of an upper bound of the variance of the random effects. An efficient maximization algorithm is presented in section 5. Section 6 presents a simulation study. Section 7 presents the analysis of a clinical trial. We conclude in section 8. \section{A POPULATION DYNAMICS MODEL}\label{DynMod} \subsection{A general model for the system}\label{syst} The dynamics of the concentrations of virions and CD4+ T-cells (in short, CD4) in different stages (represented by ${\mbox{\boldmath $X$}}^{i}(t)$) can be described by an ODE system. We allow the values of the parameters to vary between subjects; thus we consider a population model, as in Guedj, Thi\'ebaut and Commenges (2007). For subject $i$ with $i=1,...n$, this can be written: \begin{equation}\label{sys2} \left \lbrace \begin{array}{l} \frac{d{\mbox{\boldmath $X$}^{i}}(t)}{dt} = f({\mbox{\boldmath $X$}}^{i}(t), \mbox{\boldmath $\xi$}^{i}) \\ {\mbox{\boldmath $X$}} ^{i}\small{} (0)=h(\mbox{\boldmath $\xi$}^{i}) \end{array} \right. \end{equation} where $\mbox{\boldmath $X$}^{i} (t)= (X^{i} _{1}(t),...,X^{i} _{K}(t))'$ is the vector of the $K$ state variables (or components); $\mbox{\boldmath $\xi$}^{i}=(\xi_{1}^{i},...,\xi_{p}^{i})$ is a vector of $p$ individual parameters which appear naturally in the ODE system and have generally a biological interpretation. Similarly to generalized (mixed) linear models, we introduce a link function which relates $\mbox{\boldmath $\xi$}^{i}$ to a linear model involving explanatory variables and random effects: \begin{equation}\label{statist} \Psi _l (\xi_{l}^{i})=\tilde \xi_{l}^{i}=\left \{ \begin{array}{ll} \phi_l + b^i_l+\mbox{\boldmath $z_l ^{i}$}(t) \mbox{\boldmath $\beta_l$} ,& ~~ l=1,\ldots, R,\\ \phi_l +\mbox{\boldmath $z_l ^{i}$}(t) \mbox{\boldmath $\beta_l$} ,&~~ l=R+1,\ldots, p, \end{array} \right. \end{equation} where $\phi_l$ is the intercept, $\mbox{\boldmath $z_l ^{i}$}(t)$ are vectors of explanatory variables associated with the fixed effects of the $l$th biological parameter; these explanatory variables may be time-dependent, in which case the ODE system has time-dependent parameters. The $\mbox{\boldmath $\beta_l$}$'s are vectors of regression coefficients; $b_{i}=(b^i_1,\ldots,b^i_R)$ is the individual vector of random effects. We assume $b_{i} \sim \mathcal{N}(0,{\mbox{\boldmath $\Sigma$}})$ with $\Sigma$ diagonal with diagonal elements $\tau_l^2$. More general models could of course be considered. \subsection{Model for the observations} Let $Y_{ijm}$ denote the $j$th measurement of the $m$th observable component for subject $i$ at time $t_{ijm}$; we assume that: \begin{equation}\label{observation}Y_{ijm}= g_{m}(\mbox{\boldmath $X$}^i(t_{ijm})) + \epsilon_{ijm}, ~~~ i=1,...,n,~~~ j=1,...,n_{im},~~~\end{equation} for $~ m=1,...,M$, where $g_{m}(.)$ are known functions and where the $\epsilon_{ijm}$ are independent Gaussian variables with zero mean and variances $\sigma_{m}^{2}$. {The $\epsilon_{ijm}$'s are supposed independent because they represent measurement errors. The model for the observations may be complicated by the detection limits of assays leading to left-censored observations $Y_{ijm}$. \subsection{A particular model for HIV dynamics}\label{standard} For illustrating the proposed method we present a version of a rather standard model for the HIV dynamics model, close to that used by Nowak and Bangham (1996): \begin{eqnarray*}\label{ode} \frac{dT^i}{dt}&=& \lambda^i - \gamma^i T^iV^i - \mu^i_{T}T^i \\ \frac{dT^{*i}}{dt}&=& \gamma^i T^iV^i - \mu^i _{T^{*}} T^{*i} \\ \frac{dV^i}{dt}& =& \pi^i T^{*i} - \mu^i _{V}V^i \\ \end{eqnarray*} where $T^i$, $T^{*i}$ represent the concentrations (implicitly depending on $t$) of non-infected and infected CD4 respectively, and $V^i$ stands for the concentration of virus. Here the components of $\mbox{\boldmath $\xi$}^{i}_l=(\lambda^i, \gamma^i, \mu_T^i, \mu_{T^*}^i, \pi^i,\mu_V^i)$ represent rates of events such as production of new cells or particles, rates of infection after meeting between different particles. As for the $\Psi_l(.)$, we will take the natural log-transform for all the parameters: the natural log-transform can be justified if we think of the parameters as expectations of Poisson variables and has the advantage that the standard deviations of the transformed parameters may be interpreted as coefficients of variations of the estimators of the natural parameters. For the simulations, in the link equation (\ref{statist}) we will take random effects for $\lambda^i$, $\pi^i$ and $\mu^i _{T^{*}}$ (so $R=3$) with the $b^i_l$ being normal and independent with variances $\tau^2_{\lambda}$, $\tau^2_{\pi}$, $\tau^2_{\mu_{T^*}}$, respectively. For all parameters except for $\gamma^i$ we will take no explanatory variable. The effect of the treatment will be modeled as modifying $\gamma^i$ according to the equation: $$\tilde \gamma^i=\log \gamma^i=\gamma_0+\beta_1z^i_1(t)+\beta_2z^i_2(t),$$ where $z^i_1(t)$ and $z^i_2(t)$ are treatment indicators. The treatment may change with time; here we will suppose that they are fixed for $t\ge 0$ but take the value $0$ for $t<0$. We assume that at $t=0$ the patients are at the equilibrium of the system with $z^i_1(t)=z^i_2(t)=0$ and this gives important information. As for the observation equation (\ref{observation}) we will take in the simulations: \begin{eqnarray*}\label{Obs} Y_{ij1}&= & \log_{10} V^{i}(t_{ij1}) + \epsilon_{ij1}\\ Y_{ij2}&= & [T^i(t_{ij2})+T^{*i}(t_{ij2})]^{1/4} + \epsilon_{ij2} \\ Y_{ij3}&= & [T^{*i}(t_{ij3})]^{1/4} + \epsilon_{ij3} \end{eqnarray*} \section{THE HIERARCHICAL OR PENALIZED LIKELIHOOD} \subsection{Asymptotic Distribution of the MHLE} Let us consider the following model: conditionally on $b_i$, $Y_i$ has a density $f_Y(.;\theta, b_i)$, where $\theta$ is a vector of fixed parameters of dimension $q$ ($\theta \in \Theta \subset \Re ^q$) and $b_i$ are random effects (or parameters) of dimension $R$. The $(Y_i,b_i)$ are independently identically distributed (iid). Typically $Y_i$ is multivariate of dimension $n_i$. We assume that the $b_i$ have density $f_b(.; \mbox{\boldmath $\tau$})$ with zero expectation and where $\mbox{\boldmath $\tau$}$ is a vector of parameters. We denote by $P^*$ the true probability and $\theta^*$ and $\mbox{\boldmath $\tau$}^*$ the parameter values which specify the distribution of the observed $Y_i$. Typically $Y_i$ is (at least partially) observed while $b_i$ is not. Estimators of both $\theta$ and $\mbox{\boldmath $b$}=(b_1,\ldots,b_n)$ are defined as maximizing the (normalized) extended loglikelihood, called here (by abuse of language) h-loglikelihood: $$\hl{\mbox{\boldmath $b$},\mbox{\boldmath $\tau$}}= L^{\theta, \mbox{\boldmath $b$}}_{_n}-\frac{1}{n}\sum _{i=1}^{n} J(b_i; \mbox{\boldmath $\tau$}),$$ where $L^{\theta, \mbox{\boldmath $b$}}_{n}$ is the loglikelihood (normalized by $\frac{1}{n}$) for the observation conditional on $\mbox{\boldmath $b$}$, and $J(b_i; \mbox{\boldmath $\tau$})=-\log f_b(b_i; \mbox{\boldmath $\tau$})$. We denote by $(\hat \theta^{\mbox{\boldmath $\tau$}},\mbox{\boldmath $\hat b$} ^{\mbox{\boldmath $\tau$}})$ the values which maximize $\hl{\mbox{\boldmath $b$},\mbox{\boldmath $\tau$}}$ for given $\mbox{\boldmath $\tau$}$; $\hat \theta^{\mbox{\boldmath $\tau$}}$ will be called the MHLE of the parameters $\theta$. We have $\hl{\mbox{\boldmath $b$},\mbox{\boldmath $\tau$}}=\frac{1}{n}\sum _{i=1}^n {\rm hl}(Y_i;\theta,b_i, \mbox{\boldmath $\tau$})$ with ${\rm hl}(Y_i;\theta,b_i,\mbox{\boldmath $\tau$})={l(Y_i;\theta,b_i)}-J(b_i; \mbox{\boldmath $\tau$})$, where ${l(Y_i;\theta,b_i)}$ is the loglikelihood for subject $i$ conditional on $b_i$. For simpler notation we will not always make the dependence in $\mbox{\boldmath $\tau$}$ explicit and will write for instance $\hl{\mbox{\boldmath $b$}}$ for $\hl{\mbox{\boldmath $b$},\mbox{\boldmath $\tau$}}$. We shall make the additional assumptions: {\bf A1} ${l(y;\theta,b_i)}$ and $J(b_i; \mbox{\boldmath $\tau$})$ are continuous and twice-continuously differentiable functions of $\theta$ and $b_i$ for all $y$ and $\mbox{\boldmath $\tau$}$; {\bf A2} ${\rm E}_{P^*} {l(Y_i;\theta,b_i)}$ exists for all $\theta \in \Theta$. We shall derive asymptotic results for the MHLE of the fixed parameters $\theta$, which do not require that $\mbox{\boldmath $\tau$}=\mbox{\boldmath $\tau$}^*$. \begin{Lemma} Under assumptions A1 and A2 the MHLE for fixed effects are M-estimators. \end{Lemma} {\bf Proof}. Consider the profile h-loglikelihood ${\rm PHL}_n(\theta)=\hl{\mbox{\boldmath $\hat b$}(\theta)}$, where $\mbox{\boldmath $\hat b$}(\theta)={\rm argmax}_b~\hl{\mbox{\boldmath $b$}}$. $(\hat \theta^{\mbox{\boldmath $\tau$}}, \mbox{\boldmath $\hat b$}(\hat \theta^{\mbox{\boldmath $\tau$}}))$ maximizes ${\rm PHL}_n(\theta)$, thus $\hat \theta^{\mbox{\boldmath $\tau$}}$ is the profile h-likelihood estimator. Remembering that $\hl{\mbox{\boldmath $b$}}= \frac{1}{n}\sum _{i=1}^n {\rm hl}(Y_i;\theta,b_i)$, it is clear that the components of $\mbox{\boldmath $\hat b$}(\theta)$ are the $\hat b_i(\theta)={\rm argmax}_{b_i} [{\rm hl}(Y_i;\theta,b_i)]$. Thus ${\rm PHL}_n(\theta)=\hl{\mbox{\boldmath $\hat b$}(\theta)}=\frac{1}{n}\sum _{i=1}^n {\rm hl}(Y_i;\theta,\hat b_i(\theta))$. It follows that $\hat \theta^{\mbox{\boldmath $\tau$}}$ is a M-estimator because it is clear that $\hat \theta^{\mbox{\boldmath $\tau$}}$ is the maximum of $M_n(\theta)=n^{-1}\sum _{i=1}^n m_{\theta}(Y_i)$, where $m_{\theta}(y)$ is a known measurable function: here $m_{\theta}(y)={\rm hl}(y;\theta,b(y;\theta))$ where $b(y;\theta)={\rm argmax}_{b} [{\rm hl}(y;\theta,b)]$ (Van der Vaart, 1998, p 41). For the convergence result we need the additional assumption: \noindent{\bf A3} For every sufficiently small ball $U\in \Theta$, ${\rm E}_{P^*} \sup_{\theta\in U}{\rm hl}(y;\theta, b(y;\theta))< \infty$. In the convergence theorems of the MHLE we will emphasize the fact that it depends on $n$ by writing $\hat \theta^{\mbox{\boldmath $\tau$}}=\hat \theta^{\mbox{\boldmath $\tau$}}_n$. \begin{Theorem} If $\Theta$ is compact and assumption A1-A3 holds, the MHLE of fixed effects $\hat \theta^{\mbox{\boldmath $\tau$}}_n$ converges in probability toward $\theta_0^{\mbox{\boldmath $\tau$}}=argmax_{\theta}~{\rm E}_{P^*} [{\rm hl}(Y_i;\theta,\hat b_i(\theta))]$, for any $\mbox{\boldmath $\tau$}$. \end{Theorem} {\bf Proof}. By the law of large numbers $M_n(\theta) \rightarrow_p M(\theta)$ where $M(\theta)= {\rm E}_{P^*} [{\rm hl}(Y_i;\theta,\hat b_i(\theta))]$. Let us call $\theta_0^{\mbox{\boldmath $\tau$}}$ the value, that we assume unique, at which $M(\theta)$ attains its maximum. The conditions stated in the Theorem, together with the continuity assumption A1, allow us to apply Wald's consistency proof (van der Vaart, 1998, Theorem 5.14, p48). \begin{Corollary} The MHLE of the fixed parameter of the statistical model described in section \ref{DynMod} converges in probability toward $\theta_0^{\mbox{\boldmath $\tau$}}=argmax_{\theta}~{\rm E}_{P^*} [{\rm hl}(Y_i;\theta,\hat b_i(\theta))]$. \end{Corollary} {\bf Proof}. In the case of the statistical model of section \ref{DynMod} we have ${\rm hl}(Y_i;\theta,\hat b_i(\theta))=\sum _{m=1}^M[ -\frac{n_i}{2} \log \sigma^2_m - \sum _{j=1}^{n_i}\frac{(Y_{ijm}-\phi_m(t_{ijm};\theta,\hat b_i(\theta)))^2}{2\sigma^2_m}]-\sum_{r=1}^R \frac{\hat b^i_r(\theta)^2}{2 \tau^2}$, where $\phi_m(t_{ijm};\theta,b_i)=g_{m}(\mbox{\boldmath $X$}^i(t_{ijm}))$ (where $\mbox{\boldmath $X$}^i(t_{ijm})$ is the solution of the ODE system with parameters $\theta,b_i$). In case where $\sigma^2_m$ are fixed, assumption A3 is trivially satisfied because we can remove the terms involving $\sigma^2_m$ and obtain a function which is bounded by zero. If we include the $\sigma^2_m$ in the parameters that we wish to estimate, assumption A3 is satisfied since ${\rm hl}(Y_i;\theta,\hat b_i(\theta))\le \sum _{m=1}^M -\frac{n_i}{2} \log \sigma^2_m - \sum _{j=1}^{n_i}\frac{(Y_{ijm}-\phi_m(t_{ijm};\theta,\hat b_i(\theta)))^2}{2\sigma^2_m}\le \sum _{m=1}^M -\frac{n_i}{2} \log \tilde \sigma^2_m - n_i/2$, with $\tilde \sigma^2_m=\frac{\sum _{j=1}^{n_i}(Y_{ijm}-\phi_m(t_{ijm};\theta,\hat b_i(\theta)))^2}{n_i}$. It seems reasonable to conjecture that ${\rm E}_{P^*} [-\log {(Y_{ijm}-\phi_m(t_{ijm};\theta,\hat b_i(\theta)))^2}] < \infty$. We can compactify the space by taking $\Theta=\bar \Re^d$. If some parameters take an infinite value, ${\rm hl}(Y_i;\theta,\hat b_i(\theta))$ take either the value $-\infty$ or a finite value. Now the problem is to investigate whether $\theta_0^{\mbox{\boldmath $\tau$}}$ is equal or close to $\theta^*$. $M(\theta)$ can be considered as an approximation of minus the Kullback-Leibler divergence. This is obtained by replacing the expectation in $\mbox{\boldmath $b$}$ by the mode. The approximation is exact if $\phi_m$ are linear functions in $\mbox{\boldmath $b$}$ but this is not true in general. However there is a possibility of reducing the bias (see section 3.3). The asymptotic normal distribution holds for M-estimators under some regularity conditions. We make use of Theorem 5.23 of van der Vaart (1998) which only requires a Lipshitz condition on $m_{\theta}(y)={\rm hl}(y;\theta,b(y;\theta))$ that we can establish if the following assumption bearing on $u^{\theta}(y)=\dpg{{\rm hl}(y;\theta,\hat b(y;\theta))}{\theta}$ holds. We shall use $u_i^{\theta}=u^{\theta}(Y_i)$ and will give an alternative expression in formula (\ref{utet}). \noindent{\bf A4} There is a neighborhood $\Theta_0\subset \Theta$ of $\theta_0^{\mbox{\boldmath $\tau$}} $ such that the function $\dot m(y)=\sup_{\theta\in \Theta_0}u^{\theta}(y)$ has the property: ${\rm E}_{P^*}\|\dot m(Y_i)\|^2 <\infty$. \begin{Theorem} \label{asympt} Assume assumptions A1-A4 hold. Then $\sqrt n (\hat \theta_n^{\mbox{\boldmath $\tau$}}-\theta_0^{\mbox{\boldmath $\tau$}})$ is asymptotically normal with zero expectation and variance equal to $$\Sigma(\theta_0^{\mbox{\boldmath $\tau$}})= \{{\rm E}_{P^*}[H_i^{\theta_0^{\mbox{\boldmath $\tau$}}}]\}^{-1} \{{\rm E}_{P^*}[u_i^{\theta_0^{\mbox{\boldmath $\tau$}}}u_i^{\theta_0^{\mbox{\boldmath $\tau$}}T}]\} \{{\rm E}_{P^*}[H_i^{\theta_0^{\mbox{\boldmath $\tau$}}}]\}^{-1},$$ where $H_i^{\theta}=\dpg{u_i^{\theta}}{\theta}$. \end{Theorem} {\bf Proof}. The theorem follows by applying Theorem 5.23 of van der Vaart (1998). In this theorem, the main condition is that there exists a measurable function $\dot m$ with ${\rm E}_{P^*} \dot m^2 < \infty$ such that for every $\theta_1$ and $\theta_2$ in a neighborhood $\Theta_0$ of $\theta_0 $ we have: \begin {equation} \label{5.23} |m_{\theta_1}(y)-m_{\theta_2}(y)| \le \dot m(y) \|\theta_1-\theta_2\|.\end{equation} A Taylor series expansion gives: $m_{\theta_1}-m_{\theta_2}= (\theta_1-\theta_2)^T\dpg{m_{\theta}}{\theta}(\tilde \theta)$, where $\tilde \theta \in \Theta_0$. This yields: $$|m_{\theta_1}-m_{\theta_2}| \le \|\dpg{m_{\theta}}{\theta}(\tilde \theta)\| \|\theta_1-\theta_2\|\le \sup_{\theta\in \Theta_0} \|\dpg{m_{\theta}}{\theta}(\theta)\| \|\theta_1-\theta_2\|.$$ Then assumption A4 allows us applying the Theorem 5.23 of van der Vaart (1998). For applying Theorem \ref{asympt}, it remains to compute the first and second derivatives of $m_{\theta}(Y_i)$ in terms of derivatives of the likelihood conditional on the random effects. We write $l_i(\theta,\hat b_i(\theta))=l(Y_i;\theta,\hat b_i(\theta))$. Let us call $\dpg{l}{x}$ (resp.$\dpg{l}{z}$) the derivatives of $l_i(.,.)$ wrt the first (resp. the second) argument and $\dpg{J}{z}$ the derivative of $J(.)$ wrt its argument. We have $u_i^{\theta}=\dpg{l_i}{x}|_{\theta, \hat b_i(\theta)}+\dpg{l_i}{z}|_{\theta, \hat b_i(\theta)}\dpg{\hat b_i}{\theta}|_{\theta}-\dpg{J}{z}|_{\hat b_i(\theta)}.$ However, because $\hat b_i(\theta)$ maximizes ${\rm hl}(Y_i; \theta,b)$ we have \begin{equation} \label{eqbhl}\dpg{l_i}{z}|_{\theta, \hat b_i(\theta)}\dpg{\hat b_i}{\theta}|_{\theta}-\dpg{J}{z}|_{\hat b_i(\theta)}=0.\end{equation} Hence we obtain that \begin{equation} \label{utet} u_i^{\theta}=\dpg{l_i}{x}|_{\theta, \hat b_i(\theta)}.\end{equation} That is $u_i^{\theta}$ is simply the derivative of the loglikelihood as if $b$ was fixed, computed in $(\theta,\hat b_i(\theta))$. Next we have $H_i^{\theta}=\dpg{u_i^{\theta}}{\theta}=\frac{\partial^2 l_i}{\partial x^2}|_{\theta, \hat b_i(\theta)} +\frac{\partial^2 l_i}{\partial z \partial x}|_{\theta, \hat b_i(\theta)} \frac{\partial \hat b_i}{\partial \theta}|_{\theta}$. Differentiating equation (\ref{eqbhl}) wrt $\theta$ we have: $$\frac{\partial^2 l_i}{\partial x \partial z}|_{\theta, \hat b_i(\theta)}+\frac{\partial^2 l_i}{\partial z^2}|_{\theta, \hat b_i(\theta)} \frac{\partial \hat b_i}{\partial \theta}|_{\theta} -\frac{\partial^2 J}{\partial z^2}|_{\hat b_i(\theta)} \frac{\partial \hat b_i}{\partial \theta}|_{\theta}=0,$$ from which we obtain: $$\frac{\partial \hat b_i}{\partial \theta}|_{\theta}= -\left [\frac{\partial^2 l_i}{\partial z^2}|_{\theta, \hat b_i(\theta)}-\frac{\partial^2 J}{\partial z^2}|_{\hat b_i(\theta)}\right ]^{-1}\frac{\partial^2 l_i}{\partial x \partial z}|_{\theta, \hat b_i(\theta)}.$$ Hence: $$H_i^{\theta}=\frac{\partial^2 l_i}{\partial x^2}|_{\theta, \hat b_i(\theta)} -\frac{\partial^2 l_i}{\partial z \partial x}|_{\theta, \hat b_i(\theta)} \left [\frac{\partial^2 l_i}{\partial z^2}|_{\theta, \hat b_i(\theta)}-\frac{\partial^2 J}{\partial z^2}|_{\hat b_i(\theta)}\right ]^{-1}\frac{\partial^2 l_i}{\partial x \partial z}|_{\theta, \hat b_i(\theta)}.$$ In practice we can plug in the estimator $\hat \theta^{\mbox{\boldmath $\tau$}}$ to obtain an estimator of $\Sigma(\theta_0^{\mbox{\boldmath $\tau$}})$ (using the continuous mapping theorem). We may also use the observed scores and Hessian. By virtue of the law of large numbers they converge toward their expectations, and again the continuous mapping theorem allows to prove consistency of the resulting estimator. \subsection{Correction of the bias}\label{bootstrap} We have shown in section 3.2 that the MHLE $\hat \theta^{\mbox{\boldmath $\tau$}}$ tends toward $\theta_0^{\mbox{\boldmath $\tau$}}$ which is in general different from $\theta^*$; thus there is an asymptotic bias $\theta_0^{\mbox{\boldmath $\tau$}}-\theta^*$. Note that the asymptotic distribution is valid for any $\mbox{\boldmath $\tau$}$, and on the other hand, $\hat \theta^{\mbox{\boldmath $\tau$}}$ is biased even for $\mbox{\boldmath $\tau$}=\mbox{\boldmath $\tau$}^*$. Thus the problem of this approach is essentially that of the bias, although a small bias may be acceptable if it goes with a small variance. We propose to partially correct the bias by parametric bootstrap (Efron and Tibshirani, 1993). Specifically, for $s=1,\ldots,S$, generate the $b_i^{s}$ from $f_b(.,\mbox{\boldmath $\tau$})$; generate $Y_i^s$ from $f_Y(.;\hat \theta^{\mbox{\boldmath $\tau$}}, b_i^{s})$; compute the MHLE $\hat \theta^{\mbox{\boldmath $\tau$}, s}$ for these data. An estimator of the bias is $S^{-1} \sum_{s=1}^S(\hat \theta^{\mbox{\boldmath $\tau$}, s}-\hat \theta^{\mbox{\boldmath $\tau$}})$. Thus the corrected estimator, called cMHLE, is $$\check \theta^{\mbox{\boldmath $\tau$} }=\hat \theta^{\mbox{\boldmath $\tau$} }-S^{-1} \sum_{s=1}^S(\hat \theta^{\mbox{\boldmath $\tau$}, s}-\hat \theta^{\mbox{\boldmath $\tau$}}).$$ This correction slightly increases the variance. The variance of $\check \theta^{\mbox{\boldmath $\tau$} }$ can be computed through the formula ${\rm var}~ {\rm E}_{P^*} (\check \theta^{\mbox{\boldmath $\tau$} }|\hat \theta ^{\mbox{\boldmath $\tau$}}) + {\rm E}_{P^*} {\rm var} (\check \theta^{\mbox{\boldmath $\tau$} }|\hat \theta ^{\mbox{\boldmath $\tau$}})$. Neglecting the bias of the MHLE in this computation we obtain: $${\rm var}~ \check \theta^{\mbox{\boldmath $\tau$}}\approx (1+S^{-1}){\rm var}~ \hat \theta^{\mbox{\boldmath $\tau$}}$$ \section{PENALTY CHOICE}\label{crossvalidation} Profile likelihood has been proposed by Therneau and Grambsch (2000) and Lee and Nelder (2001) but it has the drawback of requiring the computation of the marginal likelihood. We propose a strategy for penalty choice which avoids this computation. For any choice of $\mbox{\boldmath $\tau$}=(\tau_1,\ldots,\tau_R)$ we have that $\hat \theta^{\mbox{\boldmath $\tau$}}$ has an asymptotic normal distribution with expectation $\theta_0^{\mbox{\boldmath $\tau$}}$ and with a variance that can be estimated. We propose to take a reasonable upper bound of $\mbox{\boldmath $\tau$}$, that is, the value $\mbox{\boldmath $\tau$}^u=(\tau^u,\ldots,\tau^u)$ where $\tau^u$ is considered as an approximate upper bound for the $\tau^*_i$. First, note that since we are working with natural logarithms of the biological parameters, the $\tau_i$ may be interpreted as coefficients of variation of these parameters. It seems reasonable (and is in agreement with the literature) to expect coefficients of variations of parameters such as rate of production of new lymphocytes ($\lambda$) or death rate of uninfected lymphocytes ($\mu_T$) are not very large, that is no more than $0.3$. \section{MAXIMIZATION ALGORITHM} Newton-like algorithms use an approximation of the Hessian of the function to maximize. Since there are many parameters, this matrix can be very large. For instance in our application $q=7$, $R=3$, $n=100$, so the number of parameters is $q+nR= 307$. In complex problems, both gradient and Hessian have to be computed numerically. Particular care must be spent to compute the Hessian both economically and precisely. The algorithm we propose is an adaptation of the Marquardt algorithm (Marquardt, 1963), taking advantage of the special structure of the Hessian in our problem. We draw two consequences of this special structure: (i) there are many terms which are equal to zero, so we do not need to compute them; (ii) the matrix is not far from being block-diagonal. We shall first consider the particular case where the number of random and fixed effects are equal ($R=q$) and the loglikelihood of subject $i$, ${l(Y_i;\theta,b_i)}$, depends only on $\theta + b_i$. We are interested in maximizing the following function: $$ \hl{\mbox{\boldmath $b$}}=\frac{1}{n}\sum_{i=1}^n \left [{l(Y_i;\theta,b_i)}-\sum_{r=1}^R\frac{{b_r^i}^2}{2\tau^2}\right ]\cdot $$ It is useful to reparametrize in term of $a_i = \theta + b_i$. One finds $$ {\rm HL}=\frac{1}{n}\sum_{i=1}^n \left [{l}^{a_i}_{i}-\sum_{r=1}^R\frac{{(a_r^i-\theta_r)}^2}{2\tau^2} \right ]=\frac{1}{n}\sum_{i=1}^n {\rm hl}_i \cdot$$ With this parameterization the loglikelihood, ${l}^{a_i}_{i}=l(Y_i;\theta,a_i-\theta)$, which is the complex part, depends only on $a_i$ so that many derivatives of the h-loglikelihood are very simple:% \begin{equation} \label{eq1} \frac{\partial {\rm HL}}{\partial \theta_r} = \frac{1}{n} \sum_{i=1}^n \frac{a^i_r - \theta_r}{\tau^2};\end{equation} $$ \frac{\partial^2 {\rm HL}}{\partial \theta_r \partial a_{r'}^i} = \frac{\delta_{rr'}}{n \tau^2}~;~ \frac{\partial^2 {\rm HL}}{\partial a_r^i \partial a_{r'}^{i'}} = 0, \textrm{if $i \neq i'$ ; } \frac{\partial^2 {\rm HL}}{\partial \theta_r \partial \theta_{r'}} = - \frac{\delta_{rr'}}{\tau^2}, $$ where $\delta_{rr'}=1$ if and only if $r=r'$. This leads to a specific block structure of the Hessian matrix, involving blocks $A =\ddpg{{\rm HL}}{\theta}= -\frac{1}{\tau^2} I_R$ and $D=\ddp{{\rm HL}}{\theta}{a_i}= \frac{1}{n \tau^2} I_R$ (where $I_R$ is the identity matrix of dimension $R$; $D$ does not depend on $i$) and $C_i= \frac{1}{n}\ddpg{{\rm hl}_i}{a_i}$; the structure is displayed in Figure \ref{hessian1}. \begin{figure}[H] \begin{center} \includegraphics[angle=270,scale=0.4]{hess_tt_alea.jpg} \caption{\label{hessian1}Hessian matrix in the case $R=q$. $\displaystyle A = -\frac{1}{\tau^2} I_R$ and $\displaystyle D = \frac{1}{n \tau^2} I_R$, $I_R$ is the identity matrix of dimension $R$, and $C_i= \frac{1}{n}\ddpg{{\rm hl}_i}{a_i}$.} \end{center} \end{figure} Fast computation of this large $(n+1)R\times (n+1)R$ Hessian matrix is possible for two reasons: (i) only the terms of blocks $C_i$ require computation of the likelihood; (ii) for computing the terms of block $C_i$ we only need to compute the second derivatives of ${l}^{a_i}_{i}$ (and not of the whole h-likelihood). Finally there are $nR(R+1)/2$ terms to compute, each involving only one computation of the solution of the ODE system (needed for the numerical differentiation): it follows that the number of computations of the solution of ODE system does not exceed that required for the computation of the Hessian for an ordinary (without random effect) non-linear model with $q$ parameters ! The nearly diagonal structure of the Hessian led us to design the so-called ``patient-by-patient'' algorithm, decoupling the optimization between patients. Denoting by $a_i(k)$ and $\theta(k)$ the values at iteration $k$, iteration $k+1$ proceeds in two steps: Step 1: For $i=1,\ldots,n$: make one Marquardt step for optimizing the function ${l}^{a_i}_{i}-\sum_{r=1}^R\frac{{(a_r^i-\theta_r(k))}^2}{2\tau^2}$ on $a_i$; this gives $a_i(k+1)$; Step 2: compute $\theta_r(k+1) = \frac{1}{n} \sum_{i=1}^n {a}^i_r(k+1) $ (which satisfies (\ref{eq1})); go to step 1 (until convergence is reached). The patient-by-patient algorithm works very well far from the maximum when the global Marquardt algorithm is hampered by the need of a large increase of the diagonal of the Hessian. However the decoupling between patients also leads to a loss of efficiency so that close to the maximum it is less efficient than the global Marquardt algorithm. This observation led us to devise a hybrid algorithm: use the patient-by-patient algorithm until all blocks $C_i$ are definite-positive; then switch to the global Marquardt algorithm. Note that ensuring that all blocks $C_i$ be definite-positive does not imply that the Hessian is so; generally however it is not far from being the case so that the Marquardt algorithm is efficient. We now consider the case where there are $R$ fixed parameters, that we call $\alpha$, associated with a random effect; such as above the loglikelihood of subject $i$, ${l(Y_i;\theta,b_i)}$, depends only on $\alpha + b_i$ and a vector of fixed parameters $\beta$. As in the preceding case, the Hessian has a particular structure (see Figure \ref{hessian2}). It involves the blocks $A$, $D$ and $C_i$ as above, and in addition blocks $B=\ddpg{{\rm HL}}{\beta}$ and $B_i=\frac{1}{n}\ddp{{\rm hl}_i}{\beta}{a_i}$. \begin{figure}[H] \begin{center} \includegraphics[angle=270,scale=0.40]{hess_ef_ea.jpg} \end{center} \caption{\label{hessian2}Hessian matrix in fixed and random effects case where $\displaystyle A = -\frac{1}{\tau^2} I_R$, $\displaystyle D = \frac{1}{n \tau^2} I_R$ (where $I_R$ is the identity matrix of dimension $R$), $B=\ddpg{{\rm HL}}{\beta}$, $B_i= \frac{1}{n}\ddp{{\rm hl}_i}{\beta}{a_i}$ and $C_i=\frac{1}{n} \ddpg{{\rm hl}_i}{a_i}$} \end{figure}% The idea, like previously, is to deal with the case where there are non-definite positive $C_i$. For that, we use the two steps of the patient-by-patient approach which give the individual parameters ${a}_i(k+1)$ and their means ${\alpha}(k+1)$. Then keeping these values fixed, we find the other fixed parameters ${\beta}(k+1)$ by a step of Marquardt algorithm with the block $B$. As soon as all blocks $C_i$ and the block $B$ are definite positive, we switch to the global Marquardt algorithm. \section{A SIMULATION STUDY}\label{simulation} \subsection{Description of the simulation study} We did simulations from the model described in section \ref{standard}. We fixed (that is we did not estimate) the parameters $\tilde \mu_V$, $\tilde \mu_T$ and $\sigma_i$, at values which are plausible in view of the literature (taking as time unit the day and as volume unit the micro-liter): $\tilde \mu_V = 3.40;~ \tilde \mu_T = -2.20;~ \sigma_i=0.5,~ i=1,2,3$. The values for the other parameters (to be estimated), including the two treatment effects $\beta_1$ and $\beta_2$, are given in Table 2. For each replica, observations for $n=100$ subjects were generated; for each subject $n_{im}=10$ observations for the three compartments ($m=1,2,3$) were generated at times $0, 3, 6, 9, 12, 15,18,21,24 ,30$. \subsection{Efficiency of the algorithm} We did a simulation to compare the number of iterations of the global Marquardt algorithm and the hybrid algorithm. We tried the two algorithms with models including one to three random effects. The initial values were: $\tilde \lambda=5.0;~\tilde \mu_{T^*}=0;~\tilde \pi=0;~\tilde \gamma_0=-5.0;~\beta_1=-1.0;~\beta_2=-1.0$. The global Marquardt algorithm did not always converge in less than 150 iterations while the hybrid algorithm nearly always converged (see Table \ref{meaniter}); when they both converged, this was toward the same values (close to the true parameter values). We checked that when we started from different values the algorithms converged toward the same values. The hybrid algorithm is faster than the global one. In Table \ref{meaniter} we give the mean number of iterations until convergence (computed on 100 replications) for which the algorithm converged in less than 150 iterations. For instance the mean number of iterations for 3 random effects was 25 versus 71 for the hybrid versus the global algorithm. The mean time of one iteration is about the same for the two algorithms. To give an idea in terms of computation time, the hybrid algorithm took about 10 mn for the case with three random effects on a standard work station (Bi Xeon, 3.8 GHz). \begin{table} \caption{\label{meaniter}Percentage of convergence in less than 150 iterations and mean number of iterations to converge with different random effects for the Global Marquardt and Hybrid algorithm.} \begin {center} \begin{tabular}{c|cc|cc} \hline \multirow{2}{*}{Random effects} & \multicolumn{2}{c}{Global algorithm} &\multicolumn{2}{c}{Hybrid algorithm} \\ & nb iter & \% success & nb iter & \% success\\ \hline $\tilde{\lambda}$ & $35$ & $81\%$ & $11$ & $100\%$ \\ $\tilde{\lambda},~\tilde{\mu}_{T^*}$ & $54$ & $59\%$ & $17$ & $100\%$ \\ $\tilde{\lambda},~\tilde{\mu}_{T^*},~\tilde{\pi}$ & $71$ & $49\%$ & $25$ & $94\%$ \\ % \hline \end{tabular} \end{center} \end{table \subsection{Efficiency of the bias correction} We estimated the bias of the corrected $\check \theta^{\mbox{\boldmath $\tau$}}$ and uncorrected MHLE $\hat \theta^{\mbox{\boldmath $\tau$}}$ using 500 replicas of a distribution with three random effects bearing on $\tilde{\lambda},~\tilde{\mu}_{T^*},~\tilde{\pi}$. We first examine the case where $\mbox{\boldmath $\tau$} =\mbox{\boldmath $\tau$}^*=(0.2,0.2,0.2)$. The biases of the uncorrected MHLE are of order $10^{-2}$ for all parameters. The correction reduces the biases to the order of $10^{-3}$ (except for one parameter), which seems negligible. \begin{table}[H] \caption{\label{biais} Parameter values for the simulation and uncorrected and corrected biases} \begin{center} \begin{tabular}{c|c|cc|cc} \hline Parameters & True value & \multicolumn{2}{c}{Mean estimated value} &\multicolumn{2}{c}{Bias} \\ $$ & $$ & non corr &corr & non corr &corr\\ \hline $\tilde \lambda$ & $4.10$ & $4.14$ & $4.10$ & $4.03~10^{-2}$ & $1.67~10^{-3}$ \\ $\tilde \mu_{T^*} $ & $-1.60$ & $-1.54$ & $-1.60$ & $5.57~10^{-2}$ & $3.67~10^{-3}$ \\ $\tilde \pi $ & $-0.170$ & $-0.160$ & $-0.166$ & $1.01~10^{-2}$ & $3.69~10^{-3}$ \\ $\gamma_0 $ & $-3.00$ & $-2.98 $ & $-3.00$ & $1.50~10^{-2}$ & $-3.60~10^{-3}$ \\ $\beta_1 $ & $-1.10$ & $-1.08$ & $-1.10$ & $2.04~10^{-2}$ & $2.59~10^{-3}$ \\ $\beta_2$ & $-1.40$ & $-1.35$ & $-1.39$ & $4.55~10^{-2}$ & $1.11~10^{-2}$ \\ % \hline \end{tabular} \end{center} \end{table} \subsection{Property of the cMHLE} We wished to check whether the asymptotic results hold in practice. We simulated data from the standard model of section \ref{standard}. In the first simulation (case 1) we took as standard deviations of the random effects $\tau^*_{\lambda}=\tau^*_{\mu_T}=\tau^*_{\mu_{T^*}}=0.2$. In a second simulation (case 2) we took $\tau^*_{\lambda}=0.1$, $\tau^*_{\mu_T}=0.2$, $\tau^*_{\mu_{T^*}}=0.3$. We did 500 replications and computed the root mean square errors (RMSE) and coverage rate of .95 confidence intervals of the estimated fixed parameters obtained in fixing the components of $\mbox{\boldmath $\tau$}^u$ in the h-likelihood at values $\tau^u=0.1; 0.2; 0.3$. The results for the RMSE are shown in Table \ref{RMSE}. In the first case, the results tend to be better when $\tau^u=0.2$ which is closer to the $\tau^*_r$, while $\tau^u=0.3$ tends to be better than $\tau^u=0.1$. For the second case the results for $\tau^u=0.2$ and $\tau^u=0.3$ were approximately of the same quality, better than for $\tau^u=0.1$. It is striking that most of the RMSE are roughly of the same order, between $10^{-2}$ and $10^{-1}$. These RMSE can be interpreted as typical relative errors on the natural parameter; in these simulation the order of magnitude is about $5\%$. In term of coverage rates, the results (see Table \ref{coverage}) are not very good for $\tau^u=0.1$. They are satisfactory for $\tau^u=0.2$ , and even more satisfactory for $\tau^u=0.3$. This corroborates our strategy based on a reasonable upper bound $\tau^u$ of the $\tau^*_r$. \begin{table}[H] \caption{Root Mean Square Error, $100$ subjects, $500$ replications. \label{RMSE}} \begin {center} \begin{tabular}{c|cc|cc|cc} \hline $\tau^u$ & \multicolumn{2}{c}{0.1} &\multicolumn{2}{c}{0.2} & \multicolumn{2}{c}{0.3}\\ \hline Par. & case 1 & case 2 & case 1 & case 2 & case 1 & case 2 \\ \hline $\tilde \lambda$ & $5.62~10^{-2}$ & $4.47~10^{-2}$ & $3.34~10^{-2}$ & $3.94~10^{-2}$ & $4.95~10^{-2}$ & $3.76~10^{-2}$ \\ $\tilde \mu_{T^*}$&$9.10~10^{-2}$ & $7.67~10^{-2}$ & $4.64~10^{-2}$ & $6.84~10^{-2}$ & $8.50~10^{-2}$ & $7.47~10^{-2}$ \\ $\tilde \pi $ & $7.27~10^{-2}$ & $5.95~10^{-2}$ & $5.14~10^{-2}$ & $5.86~10^{-2}$ & $5.25~10^{-2}$ & $5.71~10^{-2}$ \\ $\gamma_0 $ & $2.60~10^{-1}$ & $1.88~10^{-1}$ & $1.35~10^{-2}$ & $1.56~10^{-1}$ & $1.77~10^{-1}$ & $1.52~10^{-1}$ \\ $\beta_1$ & $1.74~10^{-1}$ & $1.59~10^{-1}$ & $1.01~10^{-1}$ & $1.04~10^{-1}$ & $1.04~10^{-1}$ & $9.89~10^{-2}$ \\ $\beta_2$ & $1.67~10^{-1}$ & $1.91~10^{-1}$ & $1.02~10^{-1}$ & $1.35~10^{-1}$ & $1.06~10^{-1}$ & $9.90~10^{-2}$ \\ % \hline \end{tabular} \end{center} \end{table} \begin{table}[H] \caption{Coverage rate, $100$ subjects, $500$ replications. \label{coverage}} \begin {center} \begin{tabular}{c|cc|cc|cc} \hline $\tau^u$ & \multicolumn{2}{c}{0.1} &\multicolumn{2}{c}{0.2} & \multicolumn{2}{c}{0.3}\\ \hline Par. & case 1 & case 2 & case 1 & case 2 & case 1 & case 2 \\ \hline $\tilde \lambda$ & $90\%$ & $81\%$ & $95\%$ & $92\%$ & $89\%$ & $89\%$ \\ $\tilde \mu_{T^*}$&$94\%$ & $81\%$ & $94\%$ & $88\%$ & $93\%$ & $89\%$ \\ $\tilde \pi $ & $96\%$ & $95\%$ & $94\%$ & $92\%$ & $93\%$ & $94\%$ \\ $\gamma_0 $ & $97\%$ & $97\%$ & $98\%$ & $96\%$ & $94\%$ & $95\%$ \\ $\beta_1$ & $94\%$ & $87\%$ & $97\%$ & $93\%$ & $93\%$ & $94\%$ \\ $\beta_2$ & $85\%$ & $74\%$ & $97\%$ & $93\%$ & $92\%$ & $93\%$ \\ % \hline \end{tabular} \end{center} \end{table} \section{APPLICATION TO A CLINICAL TRIAL} As an application of the proposed method, we aimed at estimating the difference of treatment effects in a randomized clinical trial (Molina et al., 1999}). The ALBI ANRS 070 trial compared over 24 weeks the combination of zidovudine plus lamivudine (AZT+3TC) with that of stavudine plus didanosine (ddI+d4T) (a third arm alternating from one regimen to another was not considered in this paper). The inclusion criteria were CD4 $\geq$ 200 cells$/\mu L$ and HIV RNA level between 4 and 5 $log_{10}$ copies/mL within 15 days before entry into the study. The primary outcome measure {defined in the study protocol} was the antiretroviral effect as measured by the mean change in HIV RNA level between baseline and 24 weeks by use of the ultra-sensitive PCR assay with lower limit of quantification of 50 copies/mL (1.7 $log_{10}$). In the main analysis of Molina et al. (1999), HIV RNA values reported as $<$ 50 copies/mL were considered equivalent to 50 copies/mL; 51 patients were included in each treatment group. Over the 24-week period, HIV RNA level declined in the two groups, with mean (SE) decreases at the end of the study of 1.26 (0.09) $log_{10}$ copies/mL in the AZT+3TC group and 2.26 (0.11) $log_{10}$ copies/mL in the ddI+d4T group. We used the model described in section 2.3. In this application only the first two components $Y_{ij1}$ and $Y_{ij2}$ were observed. Moreover only a left-censored version of $Y_{ij1}$ was observed; this was taken into account in the likelihood as in Guedj, Thi\'ebaut and Commenges (2007). In view of less informative observations than in the simulations we fixed the values of three parameters: $\tilde \mu_{T}= -2.20$, $\tilde \mu_V =3.40$ and $\gamma_0= -3$. We put random effects on $\lambda$, $\pi$ and $\mu_T^*$, working with $\tau^u=0.3$. The estimated values of the parameters in natural logarithmic scale are displayed in Table 5. Reverting to natural parameters we find: $\hat \lambda= 56.8~ [54.1; 59.7]$; $\hat \pi=0.79~ [0.67; 0.92]$; $\hat \mu_{T^*}= 0.18~ [0.17; 0.19]$. In addition it was possible to test whether the two treatment groups differed. The relevant null hypothesis is ``$\eta = 0$'', where $\eta=\beta_2-\beta_1$. A natural test statistic is $W={\hat \eta \over \sqrt{\widehat {\rm var}~ \hat \eta}}$, where $\hat \eta=\hat \beta_2-\hat \beta_1 $ and $\widehat {\rm var} ~\hat \eta$ can easily be computed from the estimate of the asymptotic variance matrix $\Sigma$. We found $\hat \eta= 0.242$, $\widehat {\rm var} ~\hat \eta = 5.16~10^{-3}$; this gives $W=3.37$ and a p-value equal to $p=7~10^{-4}$. Thus we conclude as expected that the treatment groups differ, and more precisely that the infectivity of the virus has been reduced more drastically in the ddI+d4T than in AZT+3TC group. Baseline infectivity is multiplied by a factor estimated to $e^{\hat \beta_2}=0.25$ and $e^{\hat \beta_1}=0.32$ in the ddI+d4T than in AZT+3TC groups respectively. \begin{table}[H] \caption{Estimated parameters based on the ALBI clinical trial} \begin{center} \begin{tabular}{c|c|c|c} \hline Parameters & Uncorrected Values & Corrected Values & Confidence interval\\ \hline $\tilde \lambda$ & $4.05$ &$4.04$& $[3.99;4.09]$ \\ $\tilde \pi $ & $-0.129 $ & $-0.242$ & $[-0.401;-0.083]$ \\ $\tilde \mu_{T^*} $ & $-1.74$ & $-1.73$ & $[-1.80;-1.65]$ \\ $\beta_1$ & $-1.33$ & $-1.12$& $[-1.29;-0.957]$ \\ $\beta_2$ & $-1.53$ & $-1.37$& $[-1.56;-1.17]$ \\ $\sigma_{CD4}$ & $0.173$ &$0.168$ & $[0.151;0.185]$ \\ $\sigma_{CV}$ & $0.584$ & $0.541$ & $[0.501;0.582]$\\ % \hline \end{tabular} \end{center} \end{table} \section{CONCLUSION} We have developed a hierarchical likelihood approach for inference in an HIV dynamical model. We have obtained the asymptotic distribution of the MHLE, we have derived a procedure which makes the bias negligible and we have developed an efficient maximization algorithm. Our simulations show that the whole approach works. We have shown that it could be applied to the analysis of a real data set. Rather precise estimates of the parameters were obtained. One limitation of this approach is that some parameters must be fixed because of identifiability problems. The model itself, although it is already statistically challenging, may be too simple from a biological point of view. The development of such an approach would require richer data, for instance observing the number of infected T cells. The main advantage of this approach is that it is easy to implement and very fast as compared to the two main competing approaches, likelihood and Bayesian inference. The main limitation is that it does not attempt to estimate the variances of the random effects. In our application we already have a knowledge of the range of values of these variances. Thus the method can be used for exploring possible models while likelihood or Bayesian inference can be used when estimates of the variances of the random effects are needed.\vspace{2mm} \noindent {\bf Acknowledgments.} The authors thank the investigators of the ALBI ANRS-070 clinical trial and particularly J. M. Molina (principal investigator) and G. Ch\^ene (methodologist). \section*{REFERENCES} \noindent \setlength{\parindent}{-8mm} Beal, S.L. and Sheiner, L.B. (1982) Estimating population kinetics. {\em Critical Reviews in Biomedical Engineering}, {\bf 8}, 195-222 Davidian, M. and Giltinan, D.M. (1995) {\em Nonlinear models for repeated measurements data}, Chapman \& Hall Donnet, S. and Samson, A. (2007) Estimation of parameters in incomplete data models defined by dynamical systems. {\em Journal of Statistical Planning and Inference}, {\bf 137}, 2815-283 Bradley Efron, R.J. Tibshirani (1993) {\em An Introduction to the Bootstrap.} Chapman \& Hall Fletcher, R. (1987) {\em Practical Methods of Optimization}. John Wiley \& Sons (Chichester) Guedj, J., Thi\'ebaut, R. and Commenges, D. (2007) Maximum likelihood estimation in dynamical models of HIV. {\em Biometrics}, {\bf 63}, 1198-1206 Ho, D.D., Neumann, A.U., Perelson, A.S., Chen, W., Leonard, J.M. and Markowitz, M.(1995) Rapid turnover of plasma virions and CD4 lymphocytes in HIV-1 infection. {\em Nature}, {\bf 373}, 123-126 Huang, X., Liu, D and Wu, H. (2006) Hierarchical Bayesian methods for estimation of parameters in a longitudinal HIV dynamic system. {\em Biometrics}, {\bf 62}, 413,423 Kuhn, E. and Lavielle, M. (2005) Maximum likelihood estimation in nonlinear mixed effects models. {\em Computational Statistics \& Data Analysis}, {\bf 49}, 1020-1038 Lee, Y. and Nelder, J.A. (1996) Hierarchical Generalized Linear Models. {\em Journal of the Royal Statistical Society. Series B}, {\bf 58}, pp. 619-67 Lee, Y. and Nelder, J.A. (2001) Hierarchical generalised linear models: A synthesis of generalised linear models, random-effect models and structured dispersions. {\em Biometrika}, {\bf 88}, 987-100 Lee, Y., Nelder, J.A. and Pawitan, Y. (2006) {\em Generalized linear models with random effects}, Chapman and Hall Lindstrom, M. and Bates, D. (1990) Nonlinear mixed effects models for repeated measures data. {\em Biometrics}, {\bf 46}, 673-687 McGilchrist, C.A. and Aisbett, C.W. (1991) Regression with Frailty in Survival Analysis. {\em Biometrics}, {\bf 47}, 461-466 Marquardt, D. (1963) An algorithm for least-squares estimation of nonlinear parameters. {\em SIAM Journal of Applied Mathematics}, {\bf 11}, 431-441 Molina, J.M., Ch\^ene, G., Ferchal, F., Journot, V., Pellegrin, I., Sombardier, M. N., Rancinan, C., Cotte, L., Madelaine, I., Debord, T. and Decazes, J. M. (1999) The {ALBI} trial: A Randomized Controlled Trial Comparing Stavudine Plus Didanosine with Zidovudine Plus Lamivudine and a Regimen Alternating Both Combinations in Previously untreated Patients Infected with Human immunodeficiency Virus. {\em The Journal of Infectious Diseases}, {\bf 180}, 351-358 Noh, M. and Lee, Y. (2008) Hierarchical-likelihood approach for nonlinear mixed-effects models. {\em Computational Statistics \& Data Analysis}, {\bf 52}, 3517-3527 Nowak, M.A. and Bangham, C.R.M. (1996) Population dynamics and immune response to persistent viruses. {\em Science}, {\bf 272}, 74-79 Nowak, M.A. and May R.M. (2000) {\em Virus Dynamics: Mathematical Principles of Immunology and Virology}, Oxford University Press O'Sullivan, F. (1988) Fast computation of fully automated log-density and log-hazard estimators. {\em SIAM Journal on Scientific and Statistical Computing}, {\bf 9}, 363-379 Putter, H, Heisterkamp, S.H., Lange, J.M. and de Wolf, F. (2002) A Bayesian approach to parameter estimation in HIV dynamical models. {\em Statistics in Medicine}, {\bf 21}, 2199-2214 Ramsay, J. O., Hooker, G., Campbell, D. and Cao, J. (2007) Parameter estimation for differential equtions: a generalized smoothing approach. {\em Journal of the Royal Statistical Society: Series B}, {\bf 69}, 741-796. Therneau, T.M. and Grambsch P.M. (2000) {\em Modeling survival data: extending the Cox model}, Springer van der Vaart, A. (1998) {\em Asymptotic Statistics}, Cambridge Wu, H. (2005) Statistical methods for HIV dynamic studies in AIDS clinical trials. {\em Statistical Methods in Medical Research}, {\bf 14}, 171-192. Wu, H. and Ding, A. (1999) Population HIV-1 Dynamics in Vivo: Applicable Models and Inferential Tools for Virological Data from AIDS Clinical Trials. {\em Biometrics}, {\bf 55}, 410-418. \end{document}
\section{Introduction} Terzan~5\ \citep[stellar luminosity L$_{\rm star}$=1.5$\times$10$^5$L$_\odot$,][]{1996AJ....112.1487H} is the Galactic globular cluster (GC) with the largest population of known millisecond pulsars \citep[33 MSPs,][]{2008IAUS..246..291R}. Furthermore, a high production rate for low-mass X-ray binaries (LMXBs) is expected in the extremely dense core of Terzan~5\ \citep{2005ASPC..328..231I}. The GC is located at a distance of 5.5~kpc \citep{2007A&A...470.1043O} or 8.7~kpc \citep{2002ApJ...571..818C}, only $\sim$1.7~deg above the Galactic plane (\hbox{R.A.: 17$^{\rm h}$48$^{\rm m}$04$\fs$0}, \hbox{Dec.: -24$^\circ$46'45"}). Its core (r$_{\rm c}$) and half-mass radii (r$_{\rm h}$) are 0.18' and 0.83', respectively \citep{1996AJ....112.1487H}. GCs are expected to contain intracluster gas originating from the mass loss of evolved stars. Since GCs move through the Galactic halo medium with typical velocities of $\sim$200~km~s$^{-1}$, bow shocks should form in front of them in the direction of their proper motion \citep{1995ApJ...451..200K}. These shocks have the ability to both accelerate particles and heat the gas behind them. In this scenario, electrons accelerated at the bow shock could produce diffuse non-thermal X-ray emission because of either inverse Compton (IC) scattering on ambient photon fields \citep{1995ApJ...451..200K} or non-thermal bremsstrahlung resulting from the deflection of the electrons by inter-stellar medium (ISM) nuclei \citep[][henceforth Ok07]{2007PASJ...59..727O}. Also, diffuse thermal X-ray emission can be emitted from shock-heated material trailing behind the moving GC. For a detailed discussion of the bow-shock scenario, see Ok07. The first unresolved diffuse X-ray emission from GCs (47~Tuc, $\omega$~Cen, and M~22) was reported by \citet{1982ApJ...254L..11H} using the \emph{Einstein} observatory, which was later confirmed by \citet{1995ApJ...451..200K} with \emph{ROSAT}. Recently, using \emph{Chandra}\ data, Ok07 detected significant diffuse X-ray emission from the GCs 47~Tuc, NGC~6752, M~5, and $\omega$~Cen. They find that the diffuse source at the position of $\omega$~Cen is likely to be a background cluster of galaxies. A similar extra-Galactic nature is proposed for 47~Tuc by \citet{2009arXiv0906.3583Y}. The remaining potentially GC-associated diffuse X-ray emission, from M~5 and NGC~6752, could arise from different scenarios. The emission in M~5 features an arclike morphology and exhibits a thermal spectrum (kT$<$0.1~keV), possibly from shock-heated gas. The clumpy structure seen near NGC~6752 presents a hard non-thermal spectrum ($\Gamma\sim$~2) and a radio counterpart, maybe from non-thermal bremsstrahlung emission by shock-accelerated electrons hitting nearby gas clouds. Non-thermal emission in the X-ray band may also be associated with compact objects, either directly, as detected from isolated low-mass X-ray binary systems \citep[see for instance][for such a candidate system in Terzan~5]{2005ApJ...618..883W}, or as secondary emission from the population of high-energy particles they would generate. This was modeled for millisecond pulsars by \citet[][henceforth VJ08]{2008AIPC.1085..277V} for the synchrotron radiation mechanism. The second case may translate into larger physical scales, due to the diffusion of the energetic particles away from their source. The populations of compact objects in Terzan~5\ may provide an opportunity to test these scenarios. Apparent, extended X-ray emission from the direction of GCs might also arise from a population of faint unresolved point-like sources below the detection limit of the observing instrument. In this work we analyzed an archival \emph{Chandra}\ observation to search for a diffuse emission component above the Galactic background associated with Terzan~5 . We characterized the X-ray signal using spatially resolved spectral analyses, after a careful study of the diffuse Galactic background, and briefly discuss different scenarios for the emission. \section{X-ray analysis and results} \subsection{Observation and data preparation} To search for extended diffuse X-ray emission from Terzan~5 , we analyzed the \emph{Advanced CCD Imaging Spectrometer} \citep[ACIS,][]{2003SPIE.4851...28G} data of an archival 40~ks \emph{Chandra}\ \citep{2002PASP..114....1W} observation (ObsID 3798), which was originally performed to characterize the faint X-ray point-source population of this GC \citep[][henceforth H06]{2006ApJ...651.1098H}. Only the ACIS-S3 chip was switched on, so we were only able to search for diffuse X-ray emission at angular distances $\lesssim$4' from the cluster core. A comprehensive study of diffuse X-ray emission seen from a number of Galactic GCs was performed by Ok07. However, these authors excluded Terzan~5\ from their work because the only available \emph{Chandra}\ dataset at that time suffered from serious pile-up effects because of a bright binary outburst in the field-of-view (FoV). In this paper we analyzed data from a newer observation where such an event did not occur. For the X-ray analysis we used the CIAO software version 4.1, supported by tools from the FTOOLS package and XSPEC version 12.5.0 for spectral modeling \citep{1996ASPC..101...17A}. The \emph{event1} data were reprocessed with the latest position and energy calibration (CTI correction, v4.1.3) using bad pixel files generated by \emph{acis\_run\_hotpix}. The good-time-interval (GTI) file supplied by the standard processing, which was used by H06, screens out a $\sim$4.0~ks interval of strong background flaring at the end of the observation. To remove an additional time period of 4.3~ks with a slightly increased background level, we used the light curve in the 0.5--7.0~keV energy band after the core region of the cluster and additional bright sources were removed from the data. A screening threshold of 1.0~cts/s yielded a net exposure of 31.0~ks. We chose these stricter criteria with respect to H06 because understanding the background is crucial for analyzing faint extended sources. \subsection{Extraction regions} To detect and remove point-like X-ray sources from the event-list, we ran \emph{wavdetect} on the GTI-screened dataset in three energy bands (0.5--2.0~keV, 2.0--7.0~keV and 0.5--7.0~keV). H06 used \emph{pwdetect} for the detection within r$_{\rm h}$ and \emph{wavdetect} for outer regions. In this paper we only analyzed areas outside r$_{\rm h}$, so results should be comparable. We estimated a point-source detection limit of $\sim$2\ergcm{-15} in the 0.5--7.0~keV band. For the most part our results are compatible with the sources listed by H06, Table~2. However, the shorter exposure time compared to the analysis of H06 led to a higher point-source detection threshold. Therefore we did not detect the faintest seven sources from H06 that we introduced manually into our source list. Sources were removed from the dataset using the 3$\sigma$ radius of the point spread function. Additionally, all events within r$_{\rm h}$ were disregarded. To measure the level of diffuse X-ray emission around Terzan~5 , we extracted spectra from eight concentric annular regions centered on the cluster core with radii from 1.1' to 3.9' (Fig.~\ref{fig-map}). Each ring has a width of 0.4'. We chose rings with equal width over rings with constant area to have comparable statistical quality in the spectra since the surface brightness decreases with distance from the GC. For the spectral analysis, we chose the 1--7~keV energy band. Widening the band in either direction lead to lower signal-to-noise ratios. At lower energies an increased contribution to the signal from soft thermal Galactic diffuse emission is expected. At energies above 7--8~keV the charged particle induced background component increases significantly for instruments onboard \emph{Chandra} . The mean effective area and energy response for each spectrum was calculated by weighting the contribution from each pixel by its flux using a detector map in the same energy band. To subtract the particle induced non-X-ray background (NXB), we used a background dataset provided by the calibration database, where the detector was operated in stowed position. The background spectrum for each ring was extracted from the respective region in this background dataset. To account for the time dependence of the NXB, we scaled the background by the ratio of the source and background count rates in the 9--12~keV energy band for each spectrum \citep[as described by][]{2003ApJ...583...70M}. To produce an image of diffuse X-ray emission above the particle background from the direction of Terzan~5 , we extracted counts in the 1--7~keV energy band and refilled the excluded source regions and the region inside r$_{\rm h}$ with \emph{dmfilth} using the photon distributions from rings around the excluded areas. We subtracted the particle background using the respective image from the stowed dataset after correction for the different exposures. The resulting image was corrected for relative exposure and adaptively smoothed with \emph{asmooth}. We required a minimum significance of 3$\sigma$ for the kernel size of the smoothing algorithm. The resulting smoothing radii were a few arcminutes, so that no details smaller than that scale can be seen in the smoothed image. Figure~\ref{fig-map} shows the resulting image together with all extraction regions that we used in this work. \begin{figure} \resizebox{0.98\hsize}{!}{\includegraphics[clip=]{skymap.ps}} \caption{ Smoothed, exposure corrected and NXB subtracted \emph{Chandra}\ image of diffuse X-ray emission in the 1--7~keV band around Terzan~5 . Excluded regions from point sources and the region inside r$_{\rm h}$ (white circle labeled r$_{\rm h}$) were refilled (see text). The wings seen towards the north, east and west are only marginally significant and might be artifacts of the smoothing algorithm. The FoV for this observation is drawn as a box (black). Shown are the eight annular extraction regions (dashed green lines, numbered 1 to 8) and the four pie-shaped regions (solid red lines, labeled N, E, S and W). The color scaling is linear and chosen such that the Galactic diffuse level is saturated as white. Thus only emission above the Galactic diffuse level appears in gray scales. } \label{fig-map} \end{figure} \begin{figure} \resizebox{0.98\hsize}{!}{\begin{turn}{-90}\includegraphics[clip=]{flux_plot_annuli_pbg_1000-7000.ps}\end{turn}} \caption{Radial dependence of the observed diffuse X-ray surface flux above the particle background in the 1--7\,keV band as seen with \emph{Chandra}\ (black crosses with error bars). Stars (red) denote the infrared surface brightness profile from \citet{1995AJ....109..218T}. The solid curves (green) show the X-ray point-source distribution described by a generalized King-profile \citep{2006ApJ...651.1098H} for the two extreme cases q=1.43$\pm$0.11. All profiles are scaled to match the first diffuse X-ray data point, using an exponential fit in the case of the infrared data. The vertical (blue) and the horizontal (magenta) line denote r$_{\rm h}$ \citep{1996AJ....112.1487H} and the Galactic diffuse background level, respectively.} \label{fig-radial-flux} \end{figure} Even though a significant contribution from thermal Galactic diffuse emission is expected, the spectra from the single rings were fit well enough by an absorbed power-law model for a preliminary flux estimate. The resulting fit parameters are given in Table~\ref{tab-specparams}. We found significant diffuse excess emission above the particle background in all rings and derived the surface brightness for each region by dividing the model flux by the effective extraction area, which is the geometric ring area inside the FoV minus excluded regions and bad pixels. Due to limited statistics, we fixed the column density at a default value of \hbox{N$_{\rm H}$\ = 1\hcm{22}}. Therefore, we list the observed surface fluxes in Table~\ref{tab-specparams}, as opposed to the intrinsic fluxes we provide for all other spectra. The diffuse surface flux shows a clear radial dependence (Fig.~\ref{fig-radial-flux}), which indicates that a significant part of the excess is connected to the cluster. At distances greater than $\sim$170" from the GC core, the observed surface flux seems to reach a base level of \hbox{F$_{\rm x,surf}$}\ $\approx$ 1.5\ergcm{-18}arcsec$^{-2}$ (1--7~keV). In the following section we derive the unabsorbed surface flux for the outer region by applying a more realistic physical model. \subsection{Galactic diffuse background} Terzan~5\ is close to the Galactic plane where diffuse Galactic emission becomes an important component. However, the \emph{Chandra}\ blank-sky datasets are composed of observations towards high Galactic latitudes, which would underestimate the sky background in our case. To test whether the spectrum observed from the outer three rings is compatible with thermal Galactic diffuse emission, we used a more physically reasonable model. Similar to \citet{1997ApJ...491..638K} and \citet{2005ApJ...635..214E}, who modeled the diffuse Galactic ridge emission as observed with \emph{ASCA} and \emph{Chandra} , respectively, we describe the Galactic diffuse component using a two-temperature (2-T) non-equilibrium ionization model (NEI) \citep{1984Ap&SS..98..367M}. To improve the statistical quality, we combined the outer three (175--246") rings into a single spectrum. The spectrum was adaptively binned to a minimum of 20 excess counts per bin. As background we again used the spectrum extracted from the same region in the NXB dataset. We fitted a 2-T NEI model to the outer spectrum, freezing most of the parameters to the best-fit values from Table~8 in \citet{2005ApJ...635..214E}. We left the surface brightnesses of the two components and the N$_{\rm H}$\ free to vary to account for the difference in flux and column density between the region around Terzan~5\ and the area observed by \citet{2005ApJ...635..214E}. In addition, we allowed the Si-abundance of the soft component as a free fit-parameter, because the low-ionized Si line at $\sim$1.8~keV \citep{1997ApJ...491..638K,2005ApJ...635..214E} was otherwise underestimated. The spectrum of the outer region together with the model fit is shown in Fig.~\ref{fig-spectra} (\emph{Top}). To be able to compare our results to the analysis of \citet{2005ApJ...635..214E}, we chose an energy range of 0.7--10~keV in this specific case. The best-fit values are given in Table~\ref{tab-specparams} (\emph{Outer} region). The total intrinsic surface flux of the two components is a factor of three lower than the value for the Galactic region observed by \citet{2005ApJ...635..214E}. This relation is in good agreement with the ratio between the column densities for both regions, which is $\sim$4 \citep{1990ARAA...28..215D}. Assuming that the Galactic column density seen from a certain direction is directly related to the expected flux from a diffuse Galactic component, we conclude that at least $\sim$3/4 of the total excess above particle background observed from the outer region comes from Galactic diffuse emission. \begin{table*} \caption[]{Extraction regions and results from spectral fitting} \renewcommand{\tabcolsep}{3.5pt} \begin{center} \begin{tabular}{lllllllll} \hline\hline\noalign{\smallskip} \multicolumn{1}{l}{Region} & \multicolumn{1}{l}{Distance range$^{(1)}$} & \multicolumn{1}{l}{Angular range$^{(2)}$} & \multicolumn{1}{l}{Excess counts$^{(3)}$} & \multicolumn{1}{l}{NEI: \hbox{F$_{\rm x,surf}$}$^{(4)}$} & \multicolumn{1}{l}{NEI: kT$^{(5)}$} & \multicolumn{1}{l}{PL: \hbox{F$_{\rm x,surf}$}$^{(6)}$} & \multicolumn{1}{l}{PL: $\Gamma ^{(7)}$} & \multicolumn{1}{l}{$\chi^{\scriptscriptstyle{2}}_{\scriptscriptstyle{\nu}}$(d.o.f.)} \\ \multicolumn{1}{l}{} & \multicolumn{1}{l}{(arcsec)} & \multicolumn{1}{l}{(deg)} & \multicolumn{1}{l}{} & \multicolumn{1}{l}{\tiny ($10^{-7}$ erg cm$^{-2}$ s$^{-1}$ sr$^{-1}$)} & \multicolumn{1}{l}{(keV)} & \multicolumn{1}{l}{\tiny ($10^{-7}$ erg cm$^{-2}$ s$^{-1}$ sr$^{-1}$)} & \multicolumn{1}{l}{} & \multicolumn{1}{l}{} \\ \noalign{\smallskip}\hline\noalign{\smallskip} Ring 1 & 55--79 & 0--360 & 195.2$\pm$26.5 & -- & -- & 1.01$\pm$0.25 & 1.8$\pm$0.4 & --/1.3(15) \\ Ring 2 & 79--103 & 0--360 & 278.6$\pm$31.1 & -- & -- & 0.91$\pm$0.17 & 2.1$\pm$0.5 & --/0.7(26) \\ Ring 3 & 103--126 & 0--360 & 274.8$\pm$33.3 & -- & -- & 0.64$\pm$0.14 & 2.2$\pm$0.6 & --/1.2(27) \\ Ring 4 & 126--150 & 0--360 & 259.9$\pm$34.2 & -- & -- & 0.47$\pm$0.09 & 3.0$\pm$0.8 & --/0.9(24) \\ Ring 5 & 150--174 & 0--360 & 276.7$\pm$35.3 & -- & -- & 0.44$\pm$0.08 & 2.9$\pm$0.7 & --/1.4(26) \\ Ring 6 & 174--198 & 0--360 & 192.7$\pm$34.7 & -- & -- & 0.30$\pm$0.10 & 2.8$\pm$0.8 & --/0.9(19) \\ Ring 7 & 198--222 & 0--360 & 227.5$\pm$36.2 & -- & -- & 0.26$\pm$0.08 & 3.6$\pm$1.2 & --/1.2(22) \\ Ring 8 & 222--246 & 0--360 & 240.3$\pm$37.5 & -- & -- & 0.31$\pm$0.07 & 3.5$\pm$1.2 & --/1.0(23) \\ \noalign{\smallskip}\hline\noalign{\smallskip} Inner & 55--174 & 0--360 & 1273.5$\pm$115.8 & 1.34$\pm$0.14 (soft) & 0.59 (soft) & 1.17$\pm$0.16 & 0.9$\pm$0.5 & 1.1(52)/ \\ & & 0--360 & & 5.36$\pm$0.80 (hard) & 5.0 (hard) & -- & -- & 1.3(50) \\ Outer & 175--246 & 0--360 & 825.3$\pm$62.5 & 0.58$\pm$0.19 (soft) & 0.59 (soft) & -- & -- & 1.2(36)/-- \\ & & 0--360 & & 2.1$\pm$0.7 (hard) & 5.0 (hard) & -- & -- & \\ \noalign{\smallskip}\hline\noalign{\smallskip} North & 60--120 & 45--135 & 191.0$\pm$26.4 & -- & -- & 1.44$\pm$0.63 & 1.5$\pm$1.0 & --/1.0(45) \\ East & 60--120 & 135--225 & 176.1$\pm$26.7 & -- & -- & 2.10$\pm$0.70 & 0.6$\pm$0.7 & --/0.9(44) \\ South & 60--120 & 225--315 & 230.1$\pm$27.0 & -- & -- & 2.46$\pm$0.72 & 0.8$\pm$0.7 & --/0.8(47) \\ West & 60--120 & 315--45 & 165.2$\pm$27.1 & -- & -- & 2.10$\pm$0.75 & 1.7$\pm$1.1 & --/1.0(45) \\ \hline\noalign{\smallskip} \end{tabular} \label{tab-specparams} \end{center} $^{(1)}$Inner and outer radii of the region. $^{(2)}$Angular range of the region. $^{(3)}$Excess counts after background subtraction. $^{(4)}$Intrinsic 0.7--10.0~keV surface flux of the two thermal components. $^{(5)}$Temperatures of the two thermal components. $^{(6)}$Surface flux (1--7~keV) resulting from an absorbed power-law fit. $^{(7)}$Spectral index resulting from an absorbed power-law fit. \end{table*} \subsection{Diffuse excess emission connected to Terzan~5 } In this section we focus on the diffuse emission observed from the inner five rings (55--175"). In addition to the radial dependence of the diffuse excess emission, Fig.~\ref{fig-radial-flux} shows the infrared surface brightness profile \citep{1995AJ....109..218T} and the X-ray point-source distribution \citep[King-profile from ][]{2006ApJ...651.1098H}. Both profiles are scaled to match the first diffuse X-ray data point, using an exponential fit in the case of the infrared data. \begin{figure} \resizebox{0.98\hsize}{!}{\begin{turn}{-90}\includegraphics[clip=]{spectrum_outer_8-10.ps}\end{turn}} \resizebox{0.98\hsize}{!}{\begin{turn}{-90}\includegraphics[clip=]{spectrum_inner-outer.ps}\end{turn}} \caption{ \emph{Top}: \emph{Chandra}\ spectrum from the outer annulus (175--246") with a 2-temperature non-equilibrium ionization model fit (stepped red line). All parameters are fixed to the values from \citet{2005ApJ...635..214E} except for the surface brightnesses of the two components. \emph{Bottom}: \emph{Chandra}\ spectrum from the inner annulus (55--175") with the spectrum from the outer annulus subtracted as background. The fit is an absorbed power-law model (red stepped line). The parameters for both fits are given in Table~\ref{tab-specparams}. } \label{fig-spectra} \end{figure} To investigate the nature of the diffuse excess emission observed from the inner region in more detail and to improve the statistical quality, we extracted the combined spectrum from the inner five rings (55--175"). As a first step we fitted the same 2-T NEI model to the NXB subtracted inner spectrum, binned to a minimum of 20 excess counts per bin, as was done for the outer region in the previous section. The resulting surface fluxes of the two components are listed in Table~\ref{tab-specparams} (\emph{Inner} region). Following the same argument as in the previous section, we estimate that in this case only $\sim$1/3 of the total observed emission is of diffuse Galactic origin. Together with the surface brightness showing a clear radial dependence with respect to the core of Terzan~5 , we conclude that a significant part of the observed flux is connected to the GC. As an estimate for the Galactic diffuse background component, we subtracted the outer (175--246", see previous section) from the inner spectrum. Figure~\ref{fig-spectra} (\emph{Bottom}) shows the resulting excess spectrum from the inner region, binned to a minimum of 20 excess counts per bin, together with an absorbed power-law model fit. The spectral parameters are collected in Table~\ref{tab-specparams} (\emph{Inner} region). Fitting a thermal plasma (MEKAL, $\chi^{\scriptscriptstyle{2}}_{\scriptscriptstyle{\nu}}$ = 1.4(50)) or a thermal bremsstrahlung model (BREMSS, $\chi^{\scriptscriptstyle{2}}_{\scriptscriptstyle{\nu}}$ = 1.4(50)) to the spectrum gave temperatures kT$>$17~keV and kT$>$25~keV, respectively. The most prominent feature of the Galactic thermal emission observed from the outer region is an emission line, centered on $\sim$1.8~keV. Introducing a Gaussian line at that energy or an additional thermal component to the excess spectrum from the inner annulus does not significantly improve the fit ($\chi^{\scriptscriptstyle{2}}_{\scriptscriptstyle{\nu}}$ = 1.2). Therefore, we conclude that the spectrum from the outer region describes the Galactic diffuse background component sufficiently. The total unabsorbed diffuse excess flux in the 1--7~keV band measured from the inner region above the Galactic background is F$_{\rm X}$~=~(5.5$\pm$0.8)\ergcm{-13}. Assuming a distance of 5.5~kpc the intrinsic luminosity is L$_{\rm X}$~=~(2.0$\pm$0.3)\ergs{33}. We found no indication of a variation in the spectral index with increasing radius, when subdividing the inner region into two or more sub-regions. Furthermore, there is no evidence of a directional variation in the index and flux. Table~\ref{tab-specparams} (\emph{north, east, south, and west} regions) illustrates the result for directional dependence of spectra extracted from pie-shaped regions towards the north, east, south and west with respect to the cluster center (Fig.~\ref{fig-map}). Their inner and outer radii are 60" and 120", respectively. The latter value was chosen such that the southern region is not truncated by the FoV. As background we again used the spectrum from the outer annulus. A similar result was achieved when the regions were rotated by 45~degrees. \section{Origin of the diffuse emission} The present results indicate GC-centered diffuse hard X-ray excess emission above Galactic background, which extends significantly beyond r$_{\rm h}$. In this section we briefly discuss standard thermal and non-thermal emission scenarios, leaving out more exotic possibilities, as described by, e.g., \citet{2005A&A...444L..33D}. Throughout this section we use a distance to Terzan~5\ of 5.5~kpc. The larger distance estimate (8.7~kpc) would increase the energy requirements for the models by a factor of 2.5. \subsection{Contribution from unresolved point sources} The luminosity of unresolved point sources inside r$_{\rm h}$ has been estimated by \citet{2006ApJ...651.1098H} to 8\ergs{32}. They furthermore determined the spatial surface distribution of X-ray sources in Terzan~5\ to $S(r)\propto(1 + (r/r_c)^2)^{(1-3q)/2}$ with q=1.43. From this distribution we expect to find in the 1--3 arcmin annulus only 9\% of the luminosity of unresolved point sources within r$_{\rm h}$. The expected 7\ergs{31} is much lower than the measured emission (2.0$\pm$0.3)\ergs{33}, so we conclude that the contribution from unresolved point sources is negligible. \subsection{Synchrotron radiation} One possibility for producing non-thermal emission by relativistic electrons is synchrotron radiation emission (SR), which would radiate at a frequency $\nu_{\rm syn} $=$ 120 \, (\gamma/10^4)^2 \, (\mathrm{B}/1 {\rm \mu G})$ sin$(\phi)$~MHz, where B is the strength of the magnetic field, $\gamma$ the Lorentz factor of the electrons, and $\phi$ the pitch angle between the magnetic field and the electron velocity (Ok07). Following Ok07, electrons with an energy of $\sim$10$^{14}$~eV would be needed to produce SR emission in typical Galactic magnetic fields of a few $\mu$G in the keV regime. The population of MSPs in the center of Terzan~5\ was suggested as a continuous source of such highly-energetic electrons \citep{2007MNRAS.377..920B,2009ApJ...696L..52V}. These particles propagate to the observed extension of the diffuse emission of 3' (4.8~pc) on a timescale of t$_{\rm diff}$=$3\times 10^3 \, \mathrm{B}_{\rm 1\mu G}$ years, assuming Bohm diffusion (VJ08). The cooling of electrons with energies of $\sim10^{14}$~eV in GCs is dominated by SR emission with typical cooling times of t$_{\rm cool}\approx$3$\times 10^{4}\, \mathrm{B}_{\rm 1\mu G}^{-2}$~years (VJ08). Assuming an injection spectrum with index -1, SR cooling, which depends linearly on the energy of the electrons, should change the index to -2. Since no such steepening of the spectrum is observed at the 2$\sigma$ level, t$_{\rm diff}$ $\lesssim$ t$_{\rm cool}$ is required, which would limit the magnetic field to $\sim$1 $\mu$G or it would indicate a faster diffusion of electrons. In this scenario, the population of highly energetic electrons has to radiate the observed X-ray luminosity (2\ergs{33}) on a timescale of t$_{\rm cool}$, so would require a total energy in these electrons of 1.8~$\times$ 10$^{45}\,\mathrm{B}_{\rm 1\mu G}^{-2}$~erg. Associated IC radiation in the TeV energy range should be detectable in spatial coincidence in the case of low magnetic fields $\mathrm{B} \lesssim$ few ${\rm \mu}$G, providing a test for this scenario (VJ08). \subsection{Inverse Compton emission} Non-thermal X-ray emission in GCs can also be produced by IC up-scattering of star-light photons by mildly relativistic electrons \citep{1995ApJ...451..200K}. The bow shock of the GC could provide these electrons (Ok07). The power of IC radiation $P_{\rm IC}$ emitted by a single electron is given by $P_{\rm IC} = 4/3 \,\sigma_T \,c \,\gamma^2 \,u_{\rm rad}$, where $\sigma_T$ is the Thomson cross section, $c$ is the speed of light, $\gamma$ the energy of the electrons, and $u_{\rm ph}$ the density of the target photon field \citep{1995ApJ...451..200K}. Therefore the intensity of IC emission is directly related to the energy density of the target photon field. GCs exhibit a very high stellar density in their core region, which decreases rapidly in their outskirts, resulting in a centrally peaked photon field as indicated with the distribution of the infrared surface brightness in Fig.~\ref{fig-radial-flux}. The diffuse X-ray emission presented in this paper exhibits a surface brightness profile that is roughly similar to this proxy of the density of the photon field, consistent with IC emission. The energy density $u_{\rm rad}$ of the stellar photon field scaling with $u_{\rm rad}\approx L_{\rm star}$/(4$\pi$r$^2$c) is about 40~eV/cm$^3$ and 5~eV/cm$^3$, at distances of 1' and 3', respectively, from the center of the GC. For the quantitative estimate of the total energy in electrons we adopt the model of \citet{1995ApJ...451..200K} giving 5~$\times$~10$^{49}$~($u_{\rm rad}$/~40~eV~cm$^{-3}$)~erg. In this scenario the X-ray emission should be accompanied by potentially detectable SR emission in the radio band. \subsection{Non-thermal bremsstrahlung} One additional emission process of non-thermal X-rays is non-thermal bremsstrahlung, which is produced when energetic electrons are deflected by protons and nuclei. In this scenario the flux of the emission should follow the distribution of target material. The Galactic density profile of ISM perpendicular to the Galactic plane at the relevant galactocentric distances of 1--3~kpc was constrained as a single Gaussian with a full width at half maximum of less than 200~pc and virtually no gas above 400~pc \citep{1984ApJ...283...90L}. At a distance of 5.5~kpc, Terzan~5\ would be at an offset above the disk of 160~pc and thus in an ambient gas density of a few times 0.1~cm$^{-3}$. The total energy in non-thermal electrons required for the emission of 2\ergs{33} of diffuse X-ray emission would be about 9$\times$10$^{49}$(n$_\mathrm{H}$/0.1~cm$^{-3}$)~erg if an electron energy of 20~keV is assumed (Ok07). In contrast to the asymmetric morphology detected by Ok07 for GCs in a bow-shock scenario, we did not find evidence of a non-uniform shape of the excess emission from Terzan~5 . This scenario could be tested by the presence of target material in the environment of the GC. Target material in the form of molecular clouds could be probed by carefully examining molecular emission lines that are shifted by the relative Galactic rotation velocity at the physical location of Terzan~5 . \subsection{Thermal contribution} The very high fitted temperature ($>$ 15 keV) of the diffuse X-ray emission would suggest a non-thermal origin. However, at least a thermal contribution to the total excess cannot be excluded at this point. If thermal bremsstrahlung is presumed as the emission mechanism, the temperature of the plasma can be estimated from the X-ray luminosity, the volume of the emission region, and the density of the plasma \citep{1995ApJ...451..200K}. With the radius of the emission region set to 5~pc (which corresponds to 188" at a distance of 5.5~kpc), this leads to $$\left(\frac{T}{10^7 \, \mathrm{K}} \right) \approx \left( \frac{L_{\rm X}^{\rm th}}{1.3\times 10^{33} (n/ 0.1 \, \mathrm{cm}^{-3})^2} \right)^2.$$ Assuming that the observed emission is entirely thermal, i.e., $L_{\rm X}^{\rm th}$ = $L_{\rm X}$, provides an upper bound for the temperature of the plasma. It appears that this upper bound strongly depends on the plasma density. For a typical density at the GC position of 0.1~cm$^{-3}$ (see Sect. 3.4), the upper bound on the temperature is about $10^7$~K ($\approx$ 1~keV). Only for densities lower than 0.05~cm$^{-3}$ can the temperature exceed 15~keV. To heat plasma to such high temperatures, strong shocks would be indispensable. The remnants of catastrophic events may release such strong shocks (see, e.g., \citet{2007A&A...475..883A} for a remnant of a supernova Ia and \citet{2005A&A...444L..33D} for remnants of compact binary mergers). It was proposed that Terzan~5\ may host the required mergers; e.g., \citet{2002ApJ...571..830S} for white dwarf mergers and \citet{2006NatPh...2..116G} for neutron star -- neutron star mergers. However, even supernova remnants may have difficulty producing such high temperatures, because even the hot, thermal plasma in the young remnant of the type Ia supernova remnant SN~1006 reaches a temperature of about 2~keV \citep[e.g.][]{2007A&A...475..883A}, significantly cooler than the temperature found for the thermal fit to the diffuse emission in Terzan~5 . If the diffuse X-ray emission is indeed thermal, it could also originate in principle from a background galaxy cluster that by chance coincides with the core of Terzan~5 . Since galaxy clusters with temperatures $>$10~keV are very rare \citep[e.g.][]{2002ApJ...567..716R}, such a correlation appears rather unlikely. From the available data, a contribution from thermal emission processes to the measured flux cannot be ruled out, but it is not likely to represent the dominant fraction. \section{Conclusions} We discovered diffuse hard X-ray emission from the GC Terzan~5\ with a photon index of about 1 and a peak flux density profile centered on the cluster core. The hard photon index makes a purely thermal emission scenario unlikely. Energetics would favor an SR scenario as the origin of the emission and would challenge simple IC and non-thermal Bremsstrahlung models generated by electrons accelerated by the bow shock of the GC. However, no simple model is clearly preferred to explain the observed emission, as expected from the limited statistics provided by the available X-ray dataset. Additional X-ray observations, detailed multi-wavelength informations, as well as refined modeling are needed to accurately interpret the unique properties of the diffuse X-ray radiation in Terzan~5 . \begin{acknowledgements} This research has made use of data obtained from the Chandra Data Archive and software provided by the Chandra X-ray Center (CXC) in the application packages CIAO and ChIPS. We thank the referee for the very constructive feedback. \end{acknowledgements} \bibliographystyle{aa}
\section{Introduction} Since the discovery \cite{Kamihara} of superconductivity in a family of iron-based superconductors (pnictides), there have been a large number of studies of their electronic properties that have revealed similarities between pnictides and cuprates. The theoretical prediction of a striped antiferromagnetic spin-density-wave (SDW) ground state \cite{Yildirim} was confirmed by neutron scattering \cite{Clarina}. Superconductivity is found in LaOFeAs with either hole doping \cite{Hai} or electron doping \cite{Kamihara,Chen}. First principles calculations \cite{Boeri,Singh1} find that the density of states (DOS) near the Fermi level ($E_F$) is predominantly due to Fe-$d$ orbitals. Owing to the approximate S$_4$ symmetry of the FeAs tetrahedra, these Fe-$d$ orbitals split into lower lying e$_{g}$ (d$_{x^{2}-y^{2}}$,d$_{3z^{2}-r^{2}}$) and higher lying t$_{2g}$ states (d$_{xy}$,d$_{yz}$,d$_{zx}$)\cite{Chao,Singh1}. Theoretical calculations \cite{Boeri,Singh1} suggest that superconductivity may not be caused by electron-phonon coupling. Just as in the cuprates, the antiferromagnetic instability, which is suppressed by doping, is one candidate to explain unconventional superconductivity. Angle-resolved photoemission spectroscopy (ARPES) experiments have recently been carried out in BaFe$_2$As$_2$ \cite{Evtushinsky,HDing}, a related pnictide. The superconducting gap in LaFeAsO$_{1-x}$F${_x}$ with $x \approx 10\%$ has been determined from the optical reflectance in the far-infrared region.\cite{Chen} X-ray scattering spectroscopy in the deeply inelastic (Compton) regime provides a direct probe of the correlated many-body ground state in bulk materials while avoiding the surface sensitivity of ARPES. The use of modern synchrotron sources \cite{cooper} makes it possible to investigate complex materials via the measurement of directional Compton profiles\cite{Laukkanen}. In this paper, we report first-principles computations of the 2D-projected electron momentum density (2D-EMD) and Compton profiles (CPs) in the iron-based superconductor LaOFeAs. We discuss Fermi surface (FS) images in the 2D-EMD and its anisotropy defined by subtracting a smooth isotropic function from the spectrum. Our analysis of the CPs reveals that FS features related to hole- as well as electron-pockets are more prominent in the CP when the momentum transfer vector lies along the [100] rather than the [110] direction. \section{Methods} \begin{figure*}[t] \includegraphics[width=\hsize]{fig1.eps} \caption{(Color online) (a) LDA band structure of LaOFeAs at $k_z$=0. (b) Theoretical 2D-EMD $\rho^{2d}(p_x,p_y)$ of LaOFeAs projected onto the (001) plane, normalized to $\rho^{2d}(0,0)$. (c) 2D-LCW folded momentum density. Computed hole-like (green dots) and electron-like (blue dots) Fermi surfaces are marked. Note momentum is given in (c) in units of $2\pi/a$, where $a$=7.6052 a.u.} \label{fig:EMD} \end{figure*} Our electronic structure calculations are based on the local density approximation (LDA) of density functional theory. An all-electron fully charge self-consistent semi-relativistic Korringa-Kohn-Rostoker (KKR) method is used\cite{ABkkr}. The compound LaO$_{1-x}$F$_x$FeAs has a simple tetragonal structure (space-group P4/nmm). We used the experimental lattice parameters\cite{Qiu2008} of LaO$_{0.87}$F$_{0.13}$FeAs in which no spin-density-wave order was observed in neutron-scattering. A non-spinpolarized calculation was performed and the magnetic structure was neglected. Self-consistency was obtained for $x$=0 and the effects of doping $x$ were treated within a rigid band model by shifting the Fermi energy to accommodate the proper number of electrons. The convergence of the crystal potential was approximately $10^{-4}$ Ry. The electron momentum density (EMD) $\rho(p_x,p_y,p_z)$ was calculated on a momentum mesh with step ($\Delta p_x,\Delta p_y,\Delta p_z) = (1/16a,1/16a,1/16c)2\pi$. The total number of points was $14.58\times10^{6}$ within a sphere of radius 12.8 a.u. in momentum space. The 2D-EMD $\rho^{2d}(p_x,p_y)$ was calculated as \begin{equation} \rho^{2d}(p_x,p_y)= \int \rho(p_x,p_y,p_z) dp_z \end{equation} while the Compton profile $J(p_z)$ is given by \begin{equation} J(p_z) = \int \int \rho(p_x,p_y,p_z) dp_xdp_y. \end{equation} % \section{Results and Discussions} % % % % In Fig.~\ref{fig:EMD}(a), we show the LDA band structure of LaOFeAs. For $x$=0, three bands cross the $E_F$ around the $\Gamma$ point, forming hole-like FSs [marked by green dots in (c)] while two bands cross $E_F$ around $M$($\pi$,$\pi$), forming electron-like FSs [marked by blue dots in (c)]. As electrons are added, the $\Gamma$ centered FSs shrink and completely disappear around $x$=0.13. The bands near $E_F$ are dominated by the Fe $d$ orbitals. The FeAs layers are separated by insulating LaO layers, with the result that the dispersion of these bands along $\Gamma-Z$ is small and, apart from a small $\Gamma$-centered 3D hole pocket, the FSs are quasi two-dimensional. Based on the fully three-dimensional computations, we take advantage of this quasi two-dimensionality and investigate quantities in the $k_x-k_y$ plane by integrating out the $k_z$ component. Figure \ref{fig:EMD}(b) shows a map of the theoretical 2D-EMD\cite{matsumoto2001}. This distribution can be described by an inverted bell shape with fourfold symmetry. The peak is at the zone center with tails extending over several unit cells. The dense contours around high symmetry points are signatures of the FS discontinuities. All these features are hidden behind the large inverted bell shaped signal. In order to investigate the Fermi surface topology in detail, we have employed both the 2D Lock-Crisp-West (LCW) folding\cite{lcw1973,matsumoto2001} and the 2D-EMD anisotropy. The 2D-LCW folding of the projected 2D-EMD $\rho^{2d}(p_x,p_y)$ is defined by \begin{equation} n(k_x,k_y)=\sum_{G_x,G_y}\rho^{2d}(k_x+G_x,k_y+G_y), \end{equation} where $n(k_x,k_y)$ gives the number of occupied states at the point $(k_x,k_y)$ in the first Brillouin zone by summing over all projected reciprocal lattice vectors $(G_x,G_y)$. The effect of the matrix element can be eliminated via the 2D-LCW folding process of Eq. (3), which thus provides a tool for focusing on the FS features. The theoretical 2D-LCW folding shown in Fig.~\ref{fig:EMD}(c) has been smoothed using a Gaussian function with $\Delta p=0.17$ a.u., which is typical of the resolutions available in high resolution Compton scattering experiments. The positions and sizes of the FS pockets of the undoped parent compound LaOFeAs found by our KKR band calculation are shown as green dots for hole pockets and as blue dots for electron pockets. Before the application of the aforementioned Gaussian broadening, $n(k_x,k_y)$ shows a maximum $n_{max}$=29.4 and a minimum $n_{min}$=24.5. The difference $n_{max}-n_{min}$=4.9 is consistent with five bands crossing the Fermi level in the LDA calculation. Even after including experimental resolution, the FS features are still quite visible as seen in Fig.~\ref{fig:EMD}(c). The maximum of $n(k_x,k_y)$ at $M(\pi,\pi)$ is associated with the electron pockets; the minimum at $\Gamma$ ($0$,$0$) is related to the hole pockets. \begin{figure} \includegraphics[width=\hsize]{fig2.eps} \caption{(Color online) Calculated 2D-EMD anisotropy in the parent compound LaOFeAs. Letters A and B label $\Gamma$ points for odd and even $k$-space sublattices of the hole-pockets as discussed in the text. The red lines are centered at $M$ points. The rapid changes in the momentum density along these lines are FS signatures of the electron-pockets. }\label{fig:alternating} \end{figure} Figure ~\ref{fig:alternating} shows the 2D-EMD anisotropy, found by subtracting a smooth isotropic function from the 2D-EMD. FS features show up as closely spaced contours around $\Gamma$($0$,$0$) and $M$($\pi$,$\pi$) points. The momentum density around the $\Gamma(or M)$ points in the higher zones is seen to be lower (or higher) than the average due to the presence of hole pockets (or electron pockets). The zone-to-zone variation of intensity of these features can be understood as a matrix element effect associated with the symmetry of the hole pockets at $\Gamma$ and electron pockets at $M$. For instance, the weak signal at the origin ($0$,$0$) can be understood since the bands crossing the Fermi level are predominantly $d$ orbitals, whereas only an $s$ orbital yields a significant contribution to the momentum density at the origin. Owing to interference effects, the FS features display a marked modulation from zone to zone. The Fe atoms in the unit cell are located at high symmetry positions, Fe1 (0,0,0) and Fe2 (0.5,0.5,0) (in units of lattice constants). The wavefunctions of these two Fe atoms show a constructive and destructive interference in momentum space, which can be represented by the structure factor $S_{\bf G}=1+e^{-i\pi(m+n)}$, where ${\bf G}=(m\hat{x}+n\hat{y})[2\pi/a]$ is a reciprocal lattice vector. Whereas $S_{\bf G}$ is largest when $(m+n)$ is even, $S_{\bf G}$ vanishes when $(m+n)$ is odd. Therefore, the FS features show the alternating pattern seen in Fig.~\ref{fig:alternating}. For the FS hole-pockets centered at $\Gamma$, the FS features at B for odd ($m+n$) are much weaker than those at A for even ($m+n$). Deviations from this rule are an indication of hybridization of the Fe orbitals with other orbitals. For the FS electron-pockets centered at $M$, a similar pattern is found. In Fig.~\ref{fig:alternating} we show red lines crossing the B sites. The change of the momentum density along the direction in which $\Delta m = \Delta n$ is more rapid than changes along a direction for which $\Delta m = -\Delta n$. The theoretical CPs along [100], [110] and [001] are shown in Fig.~\ref{fig:cp}(a). The CP along [001] is a smooth curve, since there are no Fermi breaks along this direction. Thus, we can use this profile for highlighting FS features from the other profiles. \begin{figure}[H] \includegraphics[width=\hsize]{fig3.eps} \caption{(Color online) (a) Directional Compton profiles(CPs) $J(p)$ along [100], [001] and [110] for $x$=0. (b)-(e) Differences $\Delta J(p)$ between various pairs of CPs for undoped ($x=0$) and doped ($x=0.15$) cases as indicated in the figures. Vertical arrows mark FS crossings with the electron pockets centered at $M$ (blue) and the hole pockets at $\Gamma$ (green) as discussed in the text. } \label{fig:cp} \end{figure} In Fig.~\ref{fig:cp}(b), we subtract the [001] from the [100] CP. The resulting periodic patterns occur at the $\Gamma$ and $M$ points and are associated with hole pockets (green arrows) and electron pockets (blue arrows) respectively. The CP has a dip within the hole pocket regions and a hump within the electron pocket regions. The FS breaks are clearly visible and should be amenable to exploration via high resolution Compton scattering experiments with statistics high enough to numerically differentiate the difference profiles. The same strategy is applied to the [110] CP in Fig.~\ref{fig:cp}(c); however, the FS features are not as clear as in the [100] direction. The main reason is that the contributions of the hole- and electron-like FSs projected on [110] overlap each other and tend to cancel out. The interference pattern acts to amplify this effect as follows. When the EMD is projected to form the CP, the projections of A and B are distinct points along [110]. For hole pockets (centered at $\Gamma$), A~(B) has strong~(weak) signals associated with FS breaks. For electron pockets [centered at $(\pi,\pi)$], the red lines in the EMD are parallel to [110] at the projection of A, while they are perpendicular to [110] at the projection of B. We notice that the FS signals are strong only along the red lines. Therefore, at the projection of A~(B), the FS signals are strong~(weak) for both hole and electron pockets. In simple terms, at A, strong hole pocket signals cancel strong electron pocket signals, while at B weak hole pocket signals cancel weak electron pocket signals. Thus [110] is not a suitable direction for studying the Fermi breaks. Next, we discuss how FS breaks disappear with electron doping. The Fermi level for $x$=0.15 shown in Fig.~1(a) at 0.075eV is obtained by a rigid band shift. At this doping level, all hole pockets around $\Gamma$ are removed. As indicated by blue arrows in Fig.~\ref{fig:cp}(d), the FS breaks associated with the FS electron pockets remain in the [100] CP. Compared to Fig.~\ref{fig:cp}(b), the dips associated with the hole pockets have completely disappeared (green arrows). In Fig.~\ref{fig:cp}(e), we subtract the [100] CP with $15$\% doping from the [100] CP with $0$\% doping. An interesting pattern of periodic maxima and minima appears around the $\Gamma$ and $M$ points which are identical with the positions of the hole and electron pockets, respectively. This may prove the most promising method of detecting FS signatures. \section{Conclusions} In conclusion, we have identified FS signatures in the momentum density of LaOFeAs, finding alternating intensity patterns in the 2D-EMD due to the symmetry of the crystal. FS signatures for both hole- and electron pockets are shown to be relatively strong in the [100] CP in comparison to the [110] CP. We thus conclude that the [100] direction is the favorable one for studying FS signatures. Our analysis further indicates that a doping-dependent experimental study should be able to determine at which doping level the hole-like FS's disappear. The present work sets a baseline for future experimental Compton scattering studies in the pnictides. \ \ \section{Acknowledgments} This work is supported by the US Department of Energy, Office of Science, Basic Energy Sciences contract DE-FG02-07ER46352, and benefited from the allocation of supercomputer time at NERSC and Northeastern University's Advanced Scientific Computation Center (ASCC). It was also sponsored by the Stichting Nationale Computer Faciliteiten (NCF) for the use of supercomputer facilities, with financial support from the Nederlandse Organisatie voor Wetenschappelijk Onderzoek (Netherlands Organization for Scientific Research).
\section{Introduction} We present a new method for finding DNA motifs. First we will describe the existing work which uses grammars to constrain the artificial evolution of programs and its application to finding patterns, particularly finding protein motifs. The Methods section~(\ref{sec:Methods}) describes how Ensembl~\cite{Ensembl:2009} DNA sequences are prepared and used. The new grammar based genetic programming (\ref{sec:GP}) is demonstrated (Section~\ref{sec:results}) by its ability to automatically find patterns early in human genes which distinguish non-protein coding genes genes from protein coding genes. \subsection{Evolving Grammars and Protein Motifs} Existing research on using grammars to constrain the artificial evolution of programs can be broadly divided in two: Grammatical Evolution \citep{oneill:2001:TEC} which uses BNF grammars and is based largely in Ireland and work in the far east using context-free grammars, tree adjoining grammars and inductive logic by Whigham, McKay and Wong. See, for example, \citet{whigham:1996:sblbGP,whigham:1999:aigp3}, \cite{Hoang:2008:IJKBIES} and~% \cite{wong:1996:aigp2}. Grammars are also used in many Bioinformatics applications, particularly dealing with sequences. Ross induced stochastic regular expressions from a number of grammars to classify proteins from their amino acid sequence \citep{ross:2001:gecco}. Regular expressions have been evolved to search for similarities between proteins, again based on their amino acid sequences~% \citep{oai:biomedcentral.com:1471-2105-8-23}. Whilst Brameier used amino acids sequences to predict the location of proteins by applying a multi-classifier~\citep{langdon:2001:eROC} linear genetic programming (GP) based approach \citep{NucPred-bioinformatics2007} (although this can be done without a grammar \citep{langdon:2005:CS}). A similar technique has also been applied to study microRNAs \citep{Brameier:2007:BMCbinf}. An interesting departure is Pappa's work which uses a grammar based GP to create application domain specific algorithms. E.g.\ \cite{pappa:2009a}, which considers prediction of protein function. While Dyrka and Nebel have used a genetic algorithm and a more powerful but also more complicated context free grammar. For example, they used a CFG when finding a meta-pattern describing protein sequences associated with zinc finger RNA binding sites \cite{Dyrka:2009:bmcBI}. Zinc finger was amongst the protein superfamilies sequence prediction tasks used by \cite{Dobson:2009:JIB}. Although Support Vector Machines can achieve high accuracy (they obtained 66.3\%) SVM models can be difficult for non-specialists to understand. Non-stochastic machine learning techniques have also been applied to DNA motifs. E.g.\ \citep{hu:2000:BI}, present a method based on decision trees, specifically C4.5. Note we are deliberately seeking intelligible motifs and so rule out approaches, such as \cite{Won:2007:bmcBI}, which evolved high performance but non-intuitive models for protein secondary structure prediction. \cite{George:2009:MBT} concisely list current computational techniques used with RNA motifs. We must be wary of over claiming. As \cite{Baird:2006:RNA} point out computational prediction is hard. Indeed they say for one problem (identification of new internal ribosome entry sites (IRES) in viral RNA) it is still not possible. Nevertheless, by concentrating on a generic tool which generates human readable motifs, of a type which are well known to Biologists, computers may still be of assistance. \section{Methods} \label{sec:Methods} \subsection{Preparation of Training Data} \label{sec:inputs} The DNA sequences for all human genes were taken from Ensembl (version~48). There are 46\,319 protein coding and 9\,836 non-coding transcripts. (Many genes have more than one transcript. There are 22\,740 coding and 9\,821 non-coding human genes.) As Table~\ref{tab:non_coding} shows most non-protein coding human genes are either pseudogenes of some sort or lead to short or micro-RNAs. \begin{table} \caption{\label{tab:non_coding} Number and type of each non-protein coding Ensembl human gene} \begin{center} \begin{tabular}{lr} pseudogene & 1516\\ snRNA & 1337\\ misc\_RNA & 1041\\ miRNA & 968\\ scRNA\_pseudogene & 843\\ snoRNA & 716\\ Mt\_tRNA\_pseudogene & 603\\ retrotransposed & 565\\ snRNA\_pseudogene & 501\\ snoRNA\_pseudogene & 486\\ rRNA\_pseudogene & 341\\ rRNA & 334\\ V\_segment & 236\\ tRNA\_pseudogene & 129\\ J\_segment & 99\\ C\_segment & 36\\ D\_segment & 32\\ Mt\_tRNA & 22\\ miRNA\_pseudogene & 21\\ misc\_RNA\_pseudogene & 7\\ Mt\_rRNA & 2\\ scRNA & 1\\ \multicolumn{1}{r}{total} & 9836 \end{tabular} \end{center} \end{table} We need to be able to check later that the automatically generated motif is general. I.e.\ it has not over fitted the examples it has seen and does not fail on new unseen examples. Therefore the protein coding and non-coding genes were randomly split in half. (Transcripts for the same gene were kept together). One half is available for training the GP and the second is never seen by GP and is reserved for demonstrating the performance of the evolved motif. The training data were then processed for use by the GP. \subsubsection{Training Data Sets for Generating DNA Motifs} Where a gene has multiple transcripts one was randomly chosen to be included in the training data. The other transcripts for the same Ensembl gene were not used for training. Figure~\ref{fig:gene_length} makes it clear that transcripts from non-coding genes tend to be shorter than those produced by protein coding genes. If the length of the transcript is known, this would be a very easy way to distinguish protein coding genes. However a classifier which simply said ``if the transcript exceeds 500 bases, the gene encodes a protein'' would tell us nothing new (even though it might be quite good at predicting). So we insist the GP seek out predictive DNA sequences. Therefore the GP is not told how long the transcript is. Instead all training data have exactly 60 bases taken from the start of the Ensembl transcript. (Transcripts less than 60 bases were not used for training). Finally duplicate sequences were removed. This gave 4639 unique non-protein coding and 11\,191 unique protein coding sequences for use as training examples. \begin{figure} \centerline{\includegraphics{graph/gene_length}} \caption{Distributions of number of bases per Ensembl human transcript. Note the length of protein coding transcripts is approximately log normally distributed. Most non-coding genes are shorter than protein coding genes. } \label{fig:gene_length} \end{figure} \subsubsection{Genetic Programming Training Set} \label{sec:training_set} To avoid unbalanced training sets, every generation all 4639 non-protein coding examples were used and 4639 coding examples were randomly chosen from the 11\,191 protein coding examples available. This is done by placing the coding examples at random in a ring. Each generation the next 4639 examples are taken from the ring. This ensures the coding examples are regularly re-used. (Each protein coding example is used once per 2.41 generations.) \subsection{Evolving DNA Motifs} \label{sec:GP} Having created training data we then use a strongly typed tree GP system \citep{poli08:fieldguide} to create an initial random population of motifs. Each generation the best 20\% are chosen and a new generation of motifs is created from them using two types of mutation (shrink and subtree) and subtree crossover \citep{poli08:fieldguide,langdon:book}. (The exact parameters are given in Table~\ref{gp.details}.) Over a number of generations the performance of the best motifs in the population improves. After 50~generations we stop the GP and take the best motif at that point and see how well it does. It is not only tested on the DNA sequences used to train it but, in order to estimate how well it does in general, it is tested also on the DNA sequences kept back (cf.\ Section~\ref{sec:inputs}). \subsubsection{Backus-Naur Form Grammar of Motifs} \label{sec:BNF} The BNF grammar is given in \cite[Figure~8, page~10]{langdon:2009:AMB}. Whilst it could be tuned to each application, this has not been necessary. In fact, we have used the same grammar for a very different task (isolating poorly performing Affymetrix cDNA probes \cite{langdon:2009:AMB}). Technical details and the reasons for its design are given in \cite{langdon:2009:AMB} and \citet{langdon:2008:CES-483}. The initial population of motifs is created by passing at random through the BNF grammar using the standard GP algorithm (ramped half-and-half \citep{poli08:fieldguide}). Although this may seem complex, \verb|gawk| (an interpreted language) is fast enough to handle populations of a million individuals. \subsubsection{Creating New Trial Motifs} \label{sec:xoverBNF} After each generation, the best 20\% of the current population are chosen to be the parents of the next generation. Each parent is allocated (on average) five children. Thus the next generation is the same size as the previous one. Children are created by either mutating high scoring parents or by recombination of two high scoring parents, cf.~\citep[Figure~2.5]{poli08:fieldguide}. In all cases the changes are made so that the resulting offspring obeys the BNF syntax rules and so are valid motifs. Therefore their performance can be estimated and (although some may perform badly) they are all still comprehensible motifs. \subsubsection{Evaluating the Motifs} Each generation each trial motif in the population is tested against the DNA sequences of the 4639 unique non-protein coding 60 base sequences available for training and 4639 protein coding 60 base unique sequences selected for use in this generation. Their performance is the sum of the number of non-coding sequences they match and the number of protein coding they do not match. However motifs which either match all or fail to match any are penalised by subtracting 4639 from their score. \section{Results} \label{sec:results} At the end of the first run, with a population of 1000 (cf.\ Table~\ref{gp.details} and Figure~\ref{fig:p1000_pops}) genetic programming produced the motif \verb'TACT|TGAT..|TA+TAT.|TA+(.CA+|T)(C|T)'. (This motif can be understood by noting the vertical bar \verb'|' indicates options. That is, if a sequence contains \verb'TACT' or \verb'TGAT..' or \verb'TA+TAT.' or \verb'TA+(.CA+|T)(C|T)' the motif is said to match it. The last two vertical bars are inside brackets \verb'()' which must be taken into account before the vertical bar they enclose. Thus the \verb'(C|T)' means either a Cytosine or a Thymine placed immediately after bases which match \verb'TA+(.CA+|T)'. The dots ``\verb'.''' mean one of the four bases must occur here. Finally \verb'A+' means a run of at least one Adenines. \begin{table \setlength{\temp}{\textwidth} \settowidth{\tempa}{Performance:} \addtolength{\temp}{-\tempa} \addtolength{\temp}{-2\tabcolsep} \caption{\label{gp.details} Strongly Typed Grammar based GP Parameters for Pseudogene and \mbox{non-coding} short RNA Prediction } \begin{tabular}{@{}lp{\temp}@{}}\hline Primitives: \rule[1ex]{0pt}{6pt} & The functions and inputs and how they are combined is defined by the BNF grammar \protect\cite[Figure~8, page~10]{langdon:2009:AMB} \\ Performance: & true positives$+$true negatives. (I.e. proportional to the area under the ROC curve or Wilcox statistic \citep{langdon:2004:ECDM}.) Less large penalty if it matches all RNA training sequences or none. \\ Selection: & (200,1000) generational, non-elitist, Population size = 1000 \\ Initial pop: & Ramped half-and-half 3:7 \\ Parameters: & 90\% subtree crossover, 5\% subtree mutation, 5\% shrink mutation. Max tree depth~17 (no tree size limit) \\ Termination: & 50 generations \\ \hline \end{tabular} \end{table} \begin{figure} \centerline{\includegraphics{graph/p1000_pops}} \caption{ Evolution of breeding population of motifs trying to locate human protein coding genes. Each generation the protein coding training cases are replaced leading to fluctuations in the measured performance. However the trend is steadily upwards. } \label{fig:p1000_pops} \end{figure} \pagebreak[4] Confusion matrices are a compact way to show the performance of prediction algorithms. They are particularly useful where there are many more examples of one class (e.g.\ protein coding) than another. An inept classifier which always said ``protein coding'' would often be correct and so have a high percentage accuracy. However it would be useless. By showing how well it does on all types of transcript a confusion matrix reveals its real performance. The matrix says how well the classifier does on each actual class (the columns). Where there are many classes, confusion matrices can also be helpful by showing where the classifier's predictions (the rows) are wrong. An good classifier will have a matrix with high values only on its leading diagonal. The following pair of confusion matrices give the evolved motif performance on its own training data (i.e.~the training data used in the last generation) and on all the data used by GP\@. Of course the actual non-coding examples are the same in the two cases. However the motif performs equally well on all the protein coding training examples as it does on the protein coding examples randomly selected for us in the last generation. (I.e.\ they are not significantly different, $\chi^2$, 1~dof.) This suggests the strategy of randomly changing training examples every generation has worked well. \begin{center} \begin{tabular}{cc} Last GP generation & GP training data\\ \begin{tabular}{l|rr|rr|} \multicolumn{1}{c}{} &\multicolumn{2}{c}{non protein } &\multicolumn{2}{c}{protein coding} \\ \cline{2-5} non protein & 3483 & (75\%) & 1403 & (30\%)\\ protein coding & 1156 & (25\%) & 3236 & (70\%)\\ \cline{2-5} \end{tabular} & \begin{tabular}{|rr|rr|} \multicolumn{2}{c}{non protein } &\multicolumn{2}{c}{protein coding} \\ \cline{1-4} 3483 & (75\%) & 3390 & (30\%)\\ 1156 & (25\%) & 7801 & (70\%)\\ \cline{1-4} \end{tabular} \end{tabular} \end{center} The next pair of confusion matrices are included for completeness. The left hand side gives the evolved motif's performance on the first 60 bases of the whole of the training data (i.e.\ including duplicates). The right hand confusion matrix refers to when the evolved pattern is applied to the whole transcript, rather than just its first 60 bases. \begin{center} \begin{tabular}{cc} All training data (60) & All training data (whole transcript) \\ \begin{tabular}{l|rr|rr|} \multicolumn{1}{c}{} &\multicolumn{2}{c}{non protein } &\multicolumn{2}{c}{protein coding} \\ \cline{2-5} non protein & 3572 & (75\%) & 3447 & (30\%)\\ protein coding & 1196 & (25\%) & 7899 & (70\%)\\ \cline{2-5} \end{tabular} & \begin{tabular}{|rr|rr|} \multicolumn{2}{c}{non protein } &\multicolumn{2}{c}{protein coding} \\ \cline{1-4} 4535 & (92\%) &11227 & (99\%)\\ 375 & (8\%) & 143 & (1\%)\\ \cline{1-4} \end{tabular} \end{tabular} \end{center} The next pair of confusion matrices contain the evolved motif's performance on all the holdout data (selecting only one transcript per gene). \begin{center} \begin{tabular}{cc} Holdout data (60) & Holdout data (whole transcript) \\ \begin{tabular}{l|rr|rr|} \multicolumn{1}{c}{} &\multicolumn{2}{c}{non protein } &\multicolumn{2}{c}{protein coding} \\ \cline{2-5} non protein & 3609 & (76\%) & 3503 & (31\%)\\ protein coding & 1159 & (24\%) & 7844 & (69\%)\\ \cline{2-5} \end{tabular} & \begin{tabular}{|rr|rr|} \multicolumn{2}{c}{non protein } &\multicolumn{2}{c}{protein coding} \\ \cline{1-4} 4529 & (92\%) &11207 & (99\%)\\ 382 & (8\%) & 163 & (1\%)\\ \cline{1-4} \end{tabular} \end{tabular} \end{center} The last pair of matrices include all transcripts for each of the hold out genes. The motif holds its performance when applied to the first 60 bases of each Ensembl transcript. However the shortness of the motif and the fact it can match the transcript at any point means the start of the transcript must be selected before using the motif otherwise performance falls. (Cf.\ the right hand of the previous two pairs of confusion matrices and the right hand of next pair.) \begin{center} \begin{tabular}{@{}cc} Holdout data (all transcripts, $\le60$) & Holdout (all transcripts, whole transcript) \\ \begin{tabular}{@{}l|rr|rr|} \multicolumn{1}{@{}l}{}&\multicolumn{2}{c}{non protein } &\multicolumn{2}{c}{protein coding} \\ \cline{2-5} non protein & 3683 & (75\%) & 6883 & (30\%)\\ protein coding & 1234 & (25\%) &16101 & (70\%)\\ \cline{2-5} \end{tabular} & \begin{tabular}{|rr|rr|} \multicolumn{2}{c}{non protein } &\multicolumn{2}{c}{protein coding} \\ \cline{1-4} 4541 & (92\%) &22778 & (99\%)\\ 376 & (8\%) & 206 & (1\%)\\ \cline{1-4} \end{tabular} \end{tabular} \end{center} Unlike many machine learning applications, there is no evidence of over fitting. Indeed the corresponding results for the holdout set are not significantly different \mbox{($\chi^2$, 3~dof)} from those on the whole training set. (Both when looking at the first 60 bases or the whole transcript). Table~\ref{tab:test_p1000} gives a break down of the evolved regular expression motif both by Ensembl human transcript type and by its components. (Note \verb!TA+(.CA+|T)(C|T)!) has been re-expressed as the union of four expressions: \verb!TA+.CA+C!, \mbox{ } \verb!TA+.CA+T!, \mbox{ } \verb!TA+TC! and \verb!TA+TT!.) The last part of the motive (i.e.\ \verb!TA+(.CA+|T)(C|T)!) typically scores more highly than the first three. However the evolved pattern succeeds at separating the non-protein coding from the protein genes by working together. It is sufficient for just one of the seven patterns to match the beginning of the gene. In many cases either several of the seven match and/or they match the DNA more than once. However the patterns are usually distinct in that, even in a gene which is matched by more than one of the 7 patterns, a part of the DNA which matches one is unlikely to also match another. Although the evolved motif has some similarity with the TATA box motif, it does not match the consensus sequence \verb'TATAAA'~\cite{Yang200752} exactly. \verb'TATAAA' occurs in the first 60 bases in 1.1\% (106) of the 9\,836 non-protein transcripts and 0.6\% (290) of the 46\,319 protein transcripts. Depending on the expected prevalence of the four bases, this is about what would be expected by chance. \section{Discussion} \label{sec:discuss} The combination of genetic programming and a BNF grammar designed for the production of intelligible patterns can be a viable way to automatically find interesting motifs in DNA and RNA sequences. The prototype system is available via \href{http://www.cs.ucl.ac.uk/staff/W.Langdon/ftp/gp-code/RE_gp.tar} {\url{ftp://cs.ucl.ac.uk/genetic/gp-code/RE\_gp.tar}}.) It has been demonstrated on a large biological DNA problem: discriminating non-protein coding from protein coding genes. The automatically generated motif \verb'TACT|TGAT..|TA+TAT.|TA+(.CA+|T)(C|T)' suggests that Thymine followed by one or more Adenine bases (particularly if the run is terminated by another Thymine or a Cytosine and Thymine) at the start of a transcript, indicates the transcript may be a short non-coding RNA sequence rather than from a protein-coding gene. \section*{Acknowledgement} This work was supported by the UK Biotechnology and Biological Sciences Research Council. under grant code BBSRC BBE0017421. \begin{landscape} \begin{table \caption{\label{tab:test_p1000} Performance of motif on first 60 bases by components and Ensembl transcript type } \footnotesize \input{test_p1000a.tex} \end{table} \end{landscape}
\section{Introduction} Magnetic field is one of the most widely utilized and powerful probes of the unconventional superconducting state. It couples to momentum and spin of the quasiparticles, via the orbital and Zeeman mechanisms respectively, and therefore can be used to probe both the momentum dependence of the order parameter and the spin structure of the Cooper pairs. In singlet superconductors both mechanisms are detrimental to superconductivity. The orbital coupling of the Cooper pair motion to the vector potential of the field leads, in type-II superconductors, to the appearance of a mixed state with partial field penetration. The Cooper pair supercurrents and the order parameter are spatially modulated and form an array of Abrikosov vortices. The upper critical field at which the second order transition into the normal state occurs is given $H_{c2}^{orb} \sim \Phi_0 / 2\pi \xi^2$, where $\Phi_0$ is the flux quantum, and $\xi$ is the superconducting coherence length~\cite{Tinkham}. For unconventional superconductors, below that field the number of the quasiparticles outside of the superconducting condensate, and, consequently, entropy-sensitive properties such as the specific heat or thermal conductivity, depend on the relative orientation of the field with respect to zeroes (nodes) or deep minima in the energy gap \cite{IVekhter:1999R,IVekhter:2001,HWon:2001,PThalmeier:2002,PMiranovic:2003,NNakai:2004,PMiranovic:2005,AVorontsov:2006,ABVorontsov:2007dos,ABVorontsov:2007kap,GRBoyd:2009}. The corresponding measurements have been extensively used to determine the symmetry of the superconducting state \cite{TPark:2003,TParkReview:2004,HAoki:2004,FYu:1995,HAubin:1997,TWatanabe:UPd2Al3,KIzawa:YNiBCkappa,KIzawa:BEDT,KIzawa:CeCoIn5,YMatsuda:2006,TSakakibara:2007,KYano:2008,KAhn:2009}. with the heavy fermion CeCoIn$_5$ and CeIrIn$_5$ two examples where a reasonably detailed comparison of theory and experiment was carried out~\cite{KIzawa:CeCoIn5,HAoki:2004,AVorontsov:2006,ABVorontsov:2007dos,ABVorontsov:2007kap}. Crucially, the anisotropic signal changes sign depending on the values of $T$ and $H$ in the superconducting phase diagram, so that either minima or maxima correspond to the direction of the field along the nodes~\cite{AVorontsov:2006,ABVorontsov:2007dos,ABVorontsov:2007kap}. Consequently, a detailed theoretical analysis is required for the interpretation of the experimental data. So far, such an analysis has only been carried out for purely orbital coupling. At the same time measurements on CeCoIn$_5$ clearly show that the second suppression mechanism, depairing due to Pauli spin polarization, is important in this compound \cite{TTayama:2002,ABianchi:2002,ABianchi:Sci08}. From the theory perspective, magnetic field aligns the spins of the unpaired electrons, while the singlet Cooper pairs cannot take advantage of the lower energy offered by a spin-polarized state. As a result, in the absence of the orbital effects, the normal state is energetically favorable above the Pauli critical field $H_P(T=0) = \Delta_s/\sqrt{2}\mu_B$ (for isotropic superconductors, see below), where $\Delta_s$ is the superconducting gap, and $\mu_B=\hbar e /2 m_e c$ is the Bohr magneton~\cite{cha62,clo62}. The normal state is reached via a second order transition at $T > T^*_P \sim 0.56 \, T_c$ and a first order transition at low temperatures $0 < T < T^*_P$. In the absence of strong anisotropy of the spin-orbit coupling, the number of unpaired electrons at $H<H_P$ does not depend on the field orientation and therefore whenever this mechanism is dominant it is not immediately obvious how efficient the field is for determination of the gap symmetry, and when exactly the inversion of the anisotropy pattern takes place. Under most circumstances the orbital suppression mechanism dominates, $H_{c2}^{orb}\ll H_P$. Pauli limiting is important in thin films and other quasi-two-dimensional materials, when the orbital coupling is inefficient~\cite{PMTedrow:1970,WWu:1995,PWAdams:1998}. In strongly layered materials even at relatively low fields the Zeeman pairbreaking effect may complicate the extraction of the nodal contribution to the specific heat~\cite{IVekhter:2001,MIchioka:2007}. Even in relatively 3D systems, if the characteristic electron velocity (Fermi velocity) is low and hence the coherence length is short, as it is in many heavy fermion and other correlated superconductors, the orbital upper critical field is high, and can be comparable to the Pauli limiting field. The question of what the vortex state anisotropies can tell us about the superconducting gap symmetry in this situation has not been explored before and is the subject of this paper. We consider a model relevant to CeCoIn$_5$, \cite{ABianchi:2002,TTayama:2002,cap04}. The situation when the orbital and Pauli pairbreaking mechanisms are comparable was considered early on by Gruenberg and Gunther~\cite{gru66} and by Maki~\cite{mak66} for an $s$-wave superconductor with spherical Fermi surface (FS). CeCoIn$_5$ has $d$-wave gap symmetry with the main $f$-electron containing Fermi surface sheet open along the $c$-axis~\cite{DHall:2001,RSettai:2001,YHaga:2001,NHarrison:2004}. Consequently, as our first task, we compute the upper critical field as a function of temperature for different orientations. We extend previous calculations~\cite{ada03,ike04,won04a} to the Fermi surface in the shape of a corrugated cylinder that gives the correct normal state resistivity anisotropy, and find the the temperature $T^*$, below which the normal to superconducting transition becomes first-order. One of our main findings is that the commonly used criterion for determining the dominant Pauli limiting regime via the Maki parameter~\cite{mak66}, $\alpha = \sqrt{2} H^{orb}_{c2}/H_P>1$ is quantitatively incorrect and depends on the symmetry of the gap and on the shape of the Fermi surface. Second, we analyse the anisotropy of the specific heat and the thermal conductivity under rotated magnetic field across the $T$-$H$ phase diagram including the effects of Pauli limiting. Since the precise details of the behavior of the anisotropy are used to infer the gap symmetry from experiment, this extension is critical for justifying $d_{x^2-y^2}$ symmetry in CeCoIn$_5$. We show that even moderately high Pauli effect has little influence on the the low temperature and low field part of the phase diagram whereas its consequences near the upper critical field are considerable. \section{Quasiclassical formulation} \subsection{Basic equations for superconductors under Zeeman and orbital field.} We use the quasiclassical formalism~\cite{eil68,lar68} at real frequencies, $\vare$, which allows to carry out calculations for arbitrary temperature and field and to self-consistently include effects of the field and impurities on the order parameter. The quasiclassical transport equation for the matrix Green's function in particle-hole and spin space, $\whg$, has the form~\cite{ale85} \bea [ (\vare + {e\over c} \vv_f(\hat{\vp}) \vA(\vR) )\, \widehat{\tau}_3 - \mu \vB \cdot \hat{\vS} - \whDelta(\vR, \hat{\vp}) - \whs_{imp}(\vR; \vare), \nonumber \\ \whg(\vR, \hat{\vp}; \vare)] + i\vv_f(\hat{\vp}) \cdot \gradR \; \whg(\vR, \hat{\vp}; \vare) = 0 \,. \hspace*{1cm} \label{eq:eilZ} \eea Here $[E_1, E_2]$ denotes a commutator, we carried out the standard~\cite{ser83} separation of the center of mass coordinate, $\vR$, and the momentum of the relative motion, $\vp$, so that the Fermi velocity depends on the position on the Fermi surface, $\vv_f(\hat{\vp})$. The orbital coupling is via the the electron charge and the vector potential, $\vA(\vR)$, and the Zeeman term, $\mu \vB \cdot \hat{\vS}$, is proportional to the electron's magnetic moment $\mu = (\mathfrak{g}/2) \mu_B$ with $\mathfrak{g}$-factor as a material-specific parameter, and the spin matrix, \be \hat{\vS} =\left( \begin{array}{cc} \vsigma & 0 \\ 0 & \vsigma^* \end{array} \right) \, \ee with $\vsigma$ the usual vector of Pauli matrices. The (retarded) Green's function in particle-hole and spin space, \be \whg^R = \left( \begin{array}{cc} g+\vg\vsigma & (f+\vf\vsigma)i\sigma_y \\ i\sigma_y (f'+\vf' \vsigma) & -g+\vg\vsigma^* \end{array} \right) \,, \ee satisfies normlization $\whg^2 = -\pi^2 \widehat{1}$. Eq.~(\ref{eq:eilZ}) is complemented by two other equations. One is the used to determine the impurity self-energy, which we treat in the self-consistent $t$-matrix approximation \cite{PJHirschfeld:1988}, $\whs_{imp}(\vR; \vare) = n_{imp} \widehat{t}(\vR; \vare)$, with \begin{equation} \widehat{t}(\vR; \vare) = u \widehat{1} + u N_0 \langle \whg (\vR, \hat{\vp}; \vare)\rangle_{\hvp} \widehat{t} \,, \end{equation} where $N_0$ is the density of states at the Fermi level. In writing this equation we assumed non-magnetic isotropic scattering with the individual impurity potential strength $u$, so that the impurity self energy does not have any momentum dependence~\cite{PJHirschfeld:1988}. The second equation is the self-consistency condition on the order parameter which relates it to the off-diagonal, in particle-hole space, component of the Green's function. Before we write it explicitly, however, it is convenient to rewrite the function $\whg$ in a different representation. Recall that we are interested in the regime $H\gg H_{c1}$, when the vortices form an Abrikosov lattice, and are considering strongly type-II superconductors. In this case the internal field $\vB$ is essentially uniform and equal to the applied field, $\vH$. The important point is that for a field which orientation is the same at all spatial points we can choose the spin quantization axis along the field $\vH=H \hat{\vz}$, and introduce for all vector quantities the notation $\vx\cdot\vsigma=\bar{x} \sigma_z$. Then the $4\times4$ Green's function, Eq.~(\ref{eq:eilZ}) consists of two blocks, \be \whg^R = \left( \begin{array}{cccc} g+\bar{g} & 0 & 0 & f+\bar{f} \\ 0 & g-\bar{g} & -f+\bar{f} & 0 \\ 0 & f'-\bar{f}' & -g+\bar{g} & 0 \\ -f'-\bar{f}' & 0 & 0 & -g-\bar{g} \end{array} \right) \,, \ee corresponding to the spin-up and spin-down components. The impurity self-energy $\whs_{imp}$ assumes the same block-form. The block structure allows us to rewrite the equations for the components of the Green's function in a simple form. We explicitly introduce the equations for the two spin components, $s = \{\uparrow(+1),\downarrow(-1)\}$, via $g_\sm{s} = g +s \bar{g}$, $f_\sm{s} = f + s \bar{f}$. These functions now satisfy independent quasiclassical equations and normalization conditions, $g_\sm{s}^2 - f_\sm{s} f'_\sm{s} = -\pi^2$, with the Zeeman energy shift ${\vare} \to {\vare} \mp \mu B$ for up and down spin respectively, see also Ref.~\onlinecite{kle04}. Now, for example, equation for the off-diagonal part of the Green's function takes the form \bea \left[-2i(\tilde{\vare}_\sm{s}-s\mu B) + \vv_f(\hat{\vp}) \left( \gradR - i \frac{2e}{\hbar c} \vA(\vR) \right) \right] \, f_\sm{s}(\vR, \hat{\vp}; \vare) \nonumber \\ = 2 \widetilde{\Delta}_\sm{s}(\vR; \vare) \, i g_\sm{s}(\vR, \hat{\vp}; \vare) \,. \label{eq:fspin} \hspace{1cm} \eea Here we introduced the shorthand notations $\tilde{\vare}_\sm{s}=\vare-\Sigma_\sm{s}$ and $\widetilde{\Delta}_\sm{s}=\Delta+\Delta_{imp,\sm{s}}$ explicitly including the diagonal ($\Sigma_s$) and off-diagonal, $\Delta_{imp,\sm{s}}$ components of $\whs_{imp}$. The spins mix only through the self-consistency equation for the singlet order parameter, \bea \Delta (\vR, \vk; \vare) &=&\int\limits_{-\infty}^{+\infty} {d\vare \over 4\pi i } \tanh {\vare\over 2T} \\ \nonumber &\times& \Big< V(\vk,\vp)\, (f_{\uparrow}(\vR,\hvp;\vare)+f_{\downarrow}(\vR,\hvp; \vare) \Big>_{\hvp} \,, \eea where $V(\vk,\vp)$ is the pairing potential. Eq.~(\ref{eq:fspin}) has the same form as the quasiclassical equation in the absence of Zeeman term, and therefore we can utilize the existing techniques to solve for each of the two spin components independently, enforcing the self-consistency for the order parameter at the final step. \subsection{Model and method of solution} We follow the approach we developed earlier~\cite{IVekhter:1999,AVorontsov:2006,ABVorontsov:2007dos,ABVorontsov:2007kap} and solve the equations using a modified Brandt-Pesch-Tewordt (BPT) approximation~\cite{BPT:1967,WPesch:1975}. We assume the existence of the Abrikosov vortex lattice and model the spatial dependence of the order parameter by $\Delta (\vR, \vp)= \Delta(\hvp) \braket{\vR}{0} = \Delta \cY(\hvp) \braket{\vR}{0}$, where $\cY(\hvp)$ is the normalized ($\langle \cY(\hvp)^2\rangle_{\hvp}=1$) basis function for the irreducible representation corresponding to the chosen gap symmetry, and normalized by a proper choice of $C_{k_y}$ coefficients the spatial vortex lattice profile, \begin{equation} \braket{\vR}{0}= \sum_{k_y} C_{k_y} {e^{ik_y\sqrt{S_f} y} \over \sqrt[4]{S_f \Lambda^2}} \widetilde\Phi_0\left( x,k_y \right)\,. \label{eq:DeltaPhi} \end{equation} The magnetic length $\Lambda^2=hc/2eB$, and \begin{equation} \widetilde\Phi_0(x,k_y)=\Phi_0\left( {x-\Lambda^2 \sqrt{S_f} k_y\over \Lambda \sqrt{S_f}} \right) \,, \label{Phi-0} \end{equation} is the ground state oscillator function, and $x$ and $y$ are in the direction normal to the field.~\cite{ABVorontsov:2007dos} Here we approximated the vortex lattice by only the superposition of the lowest oscillator wave functions, $\ket{0}$. The admixture of the higher oscillator states is small~\cite{ILukyanchuk:1987,HWon:1996} and does not substantially affect the conclusions regarding the properties in the vortex state~\cite{HWon:1996,AVorontsov:2006,ABVorontsov:2007dos,ABVorontsov:2007kap}. Hence we consider only the lowest Landau level, as reflected in Eq.~(\ref{eq:DeltaPhi}), but with properly rescaled Fermi velocity component perpendicular to the field. For a Fermi surface rotationally invariant around the $c$ axis that we consider below~\cite{ABVorontsov:2007dos} \be S_f = \left[\cos^2 \theta_H + {v_{0||}^2\over v_{0\perp}^2} \sin^2 \theta_H\right]^{1/2} \,, \ee $\theta_H$ is the angle between the field direction and the $c$-axis, $v_{0\parallel}^2 = 2 \langle \cY^2(\hvp) v^2_\parallel(p_z) \rangle_\sm{FS}$, and $v_{0\perp}^2 = 2 \langle \cY^2(\hvp) v^2_{\perp i}(p_z) \rangle_\sm{FS}$, where $v_\parallel$ is the $c$-axis component of the Fermi velocity, while $v_{\perp i}$ with $i=a,b$ is the Fermi velocity component in the $a$-$b$ plane. For the field in the basal plane $\theta_H=\pi/2$, and therefore $S_f=v_{0||}/ v_{0\perp}$. The BPT approximation consists of replacing the diagonal part of the Green's function with its spatial average, and is justified over a wide range of fields~\cite{EHBrandt:1995,TDahm:2002}. The equations for the off-diagonal components of the Green's function are solved by introducing the ladder operators as in Ref.~\onlinecite{ABVorontsov:2007dos,ABVorontsov:2007kap}, and, in conjunction with the normalization condition, for the lowest Landau level give \bea && f_s= \pi \frac{1}{\sqrt{1+P}} \frac{2\sqrt{\pi}\Lambda}{|\tilde{v}_f^\perp|} \, W \left[ \frac{2(\tilde{\vare}-s\mu B)\Lambda}{|\tilde{v}_f^\perp|} \right] \, \tilde\Delta_s \,, \nonumber \\ && g_s= \pi \frac{-i}{\sqrt{1+P}} \,, \label{eq:gs} \\ && P = -i\sqrt{\pi}\left(\frac{2\Lambda}{|\tilde{v}_f^\perp|}\right)^2 \, W^\prime \left[ \frac{2(\tilde{\vare}-s\mu B)\Lambda}{|\tilde{v}_f^\perp|} \right] \, \widetilde\Delta_s \ul{\widetilde\Delta}_s \,, \nonumber \eea with $ \widetilde\Delta_s(\hvp,\vare) = \Delta(\hvp) + \Delta_{imp,s}(\vare)$, $W(z) = \exp(-z^2) \mbox{erfc}(-iz)$, and \begin{equation} |\tilde{v}_f^\perp|=\left[\frac{v_{f,x}(\hat{\vp})^2}{{S_f}}+ v_{f,y}^2(\hat{\vp}) {S_f}\right]^{1/2}\,. \end{equation} This closed form solution is used to enforce the self-consistency on the impurity self-energy and the gap value. The approach gives the order parameter and the Green's function that can be used to determine the physical properties below. \section{Fermi surface, upper critical field, and Maki parameter} We consider a model Fermi surface that approximates the main sheet of CeCoIn$_5$ as seen by the magnetic oscillations ~\cite{RSettai:2001,YHaga:2001,HShishido:2002,NHarrison:2004}. It has the shape of an open cylinder rotationally symmetric in the $a$-$b$ plane (component of the momentum labeled $p_r$) and modulated along the $c$-axis ~\cite{AVorontsov:2006}, and is given by $p^2_f = p_r^2-r^2 p_f^2 \cos(2 a p_c/r^2 p_f)$ with $r=a=0.5$. This choice of parameters gives a moderate anisotropy between transport coefficients in the $c$-direction and $a$-$b$ plane, close to that of CeCoIn$_5$ in the normal state~\cite{AMalinowski:2005}. With this definition, the typical quasiparticle velocity in the basal plane is $v_f = p_f/m$, and that along the $c$-axis, $v_{f,c} \sim a v_f = 0.5 v_f$. We choose the $d_{x^2-y^2}$ symmetry of the order parameter and take a model separable pairing interaction $V(\vk,\vp)=V_0\cY(\hvk)\cY(\hvp)\equiv V_0\cY(\phi)\cY(\phi^\prime)$, where $\phi$ labels the azimuthal angle around the Fermi surface, and $\cY(\phi)=\sqrt{2}\cos 2\phi$. With this choice $\Delta(\vR,\hvp) = \sqrt{2} \Delta(\vR) \cos2\phi$. The dimensionless coupling constant $N_0V_0$ determines the transition temperature of the pure sample, $T_{c0}$, which is suppressed by impurities to $T_c$; we use this latter $T_c$ as a unit of energy. Similarly, the natural unit for the magnetic field is the characteristic orbital scale, $B_0 = \Phi_0/2\pi\xi_0^2$, where $\Phi_0=hc/2e$ is the flux quantum, and $\xi_0 = \hbar v_f/2\pi T_c$ is the coherence length. We measure the strength of the Zeeman term via a dimensionless parameter $Z = \mu B_0 / 2\pi T_c$,\cite{ada05} so that the Zeeman splitting of the energy levels in dimensionless units is $\mu B/(2\pi T_c)= Z \: B/B_0$. Since we expect the orbital critical field $H_{c2}^{orb}$ to be of order $B_0$, and the Pauli limit to correspond to $\mu H_P \sim \Delta_0 \sim T_c$, we find $H^{orb}_{c2} / H_P \sim Z$. We first compute the upper critical field by solving the linearized, with respect to the order parameter, quasiclassical equations. Under the combined effect of the orbital and Zeeman field the transition for strong enough Pauli term becomes first order at low temperatures, $T<T^*$. We determine $T^*$ in the clean limit by evaluating where the $\Delta^4$-term coefficient in the free energy expansion becomes negative. The general free energy expansion can be obtained from Eqs.~(\ref{eq:gs}), but it is complicated in a dirty $d$-wave superconductor\cite{AVorontsov:FFLOimp}, and since the precise location of $T^*$ is not important for this work we will not look for it here. At lower temperatures, the transition line can only be determined from the full free energy functional, which needs to also account for the possible existence of the additional modulations in the Fulde-Ferrell-Larkin-Ovchinnikov (FFLO) phase~\cite{ada03,ike04,won04a}. While there are many indications that in CeCoIn$_5$ a new phase exists in the high-field, low-temperature range~\cite{ABianchi:2003,cap04,wat04,mar05,VMitrovic:2006,YMatsuda:2007} the experiments on the vortex state anisotropy are generally carried out away from that range. Consequently, we do not consider the FFLO-like modulation here. \begin{figure}[t] \centerline{\includegraphics[height=5.5cm]{Fig1.eps}} \caption{(Color online) Illustration of the possible pitfalls in the experimental determination of orbital limit on upper critical field, $H^{orb}_{c2}$, in strongly paramagnetic superconductors. (a) in-plane field, (b) field normal to the planes. Lower (blue) solid lines denote $H_{c2}$ for the paramagnetically limiting case $Z=0.8$, and the dashed lines mark regions of the first-order transition. ``Turning off'' Zeeman effect ($Z=0$) results in true orbital $H_{c2}$, the upper (black) solid lines. Dotted lines: slope $dH_{c2}/dT$ at $T_c$ determined from the points $T/T_c=(0.9,\;1)$. Thick dot-dashed lines: upper critical field obtained using the Helfand-Werthamer (HW) result~\cite{hel66} for the given slope $dH_{c2}/dT$ at $T_c$. Thin dot-dashed lines: HW profile of $H_{c2}(T)$ for the paramagnetically limited case $Z=0.8$, resulting in incorrect extracted orbital limit, $H_{c2}^{ex}$. Vertical arrows indicate potential discrepancy between the approximately determined and realistic orbital upper critical field, see text for details. } \label{fig:hc2ex} \end{figure} Fig.~\ref{fig:hc2ex} shows the upper critical field, $H_{c2} (T)$ for a pure system for two different field orientations ($c$-axis and the in-plane along the gap maximum) and for different strength of the Zeeman splitting. Our results for $Z=0$ resemble the classical results by Helfand and Werthamer \cite{hel64,hel66} (HW) with differences in profiles attributed to anisotropic gap and non-spherical FS. The important observation is that the shape of the temperature dependence of the upper critical field changes with increasing $Z$, and that the linear region of $H_{c2}(T) \sim 1-T/T_c$ near $T_c$ rapidly shrinks, eventually leading to $H_{c2} \sim \sqrt{1-T/T_c}$ dependence characteristic of the Pauli limited field. Hence, as is shown in Fig.~\ref{fig:hc2ex} a frequently used experimental estimate of the $H^{orb}_{c2} (T=0)$ based on the slope of the measured upper critical field near $T_c$ is not reliable for paramagnetically limited superconductors: that estimate changes on approaching the Pauli limit. First, the slopes $dH_{c2}/dT$ of the purely orbital and paramagnetically limited cases, while equal asymptotically as $T\rightarrow T_c$, are different when determined within a reasonable experimental window as shown by dotted lines based on the values of $H_{c2}$ at $T/T_c=(0.9,\;1)$. We find that even for moderate paramagnetic coupling a reasonable estimate of the slope may be obtained only within a window of 1-3\% near $T_c$. Second, the $H_{c2}(T)$ profile for our chosen Fermi surface and the $d$-wave order parameter even for the purely orbital case is different from that found using the HW~\cite{hel66} result for $s$-wave superconductor with a spherical FS with the same $dH_{c2}/dT \,(T_c)$. Consequently, for $Z \sim 1$ the estimate of $H_{c2}^{orb}$ using the HW profile and an approximate slope $dH_{c2}/dT \,(T_c)$ based on points more than a few percent away from $T_c$ (thin dot-dashed lines) may significantly underestimate the magnitude of $H_{c2}^{orb}$, as indicated by the vertical arrows in Fig.~\ref{fig:hc2ex}. \begin{figure}[t] \centerline{\includegraphics[height=5.5cm]{Fig2.eps}} \caption{(Color online) Upper critical field of a pure $d_{x^2-y^2}$ superconductor with quasi-cylindrical FS for three orientations of the field. The critical value of the Maki parameter is about $\alpha^* \sim 3$ for all directions (see text). The dotted lines indicate first-order transition below $T^*$. Inset in the right-hand panel: $T^*$ as a function of the relative strength of the Pauli limiting. For purely Pauli-limited case $T^*_P \approx 0.56 T_c$. } \label{fig:hc2} \end{figure} The change in the shape of the $H_{c2}(T)$ curve with increasing paramagnetic contribution is made even more explicit in Fig.~\ref{fig:hc2}, where we also include the upper critical field for the in-plane field along the nodal direction. As can be expected in a magnetically isotropic system, increased Pauli limiting reduces the anisotropy of the upper critical field caused by both the in-plane anisotropy of the order parameter (node vs. anti-node) and the the anisotropy of the Fermi surface ($c$-axis vs. $a$-$b$ plane). Finally, in the inset of Fig.~\ref{fig:hc2} we show the onset temperature for the first order transition, $T^*$ as a function of the Pauli limiting parameter $Z$. It is instructive to recast this analysis in terms of the so-called Maki parameter, $\alpha = \sqrt{2} H^{orb}_{c2}/H_P$. It is conventionally assumed that the critical value of the Maki parameter that defines a strong paramagnetic limit is $\alpha > \alpha^* = 1$, as obtained for the onset of the first order transition in $s$-wave superconductors with a spherical Fermi surface~\cite{mak66}; this value is commonly used in the analysis of anisotropic strongly correlated materials as well. In reality the critical value $\alpha^*$ depends on details of the Fermi surface and the superconducting state. For our model $H_{c2}^{orb} \sim 1.4$ in the $a$-$b$ plane, and $\mu H_P \approx 1.11 \Delta_{d0}/\sqrt{2}$. Here $\Delta_{d0}= 0.241 (2\pi T_c)$ is the amplitude of the order parameter at $T=H=0$\cite{vor05b} and the increase of $H_P$ by factor $1.11$ compared with the similar $s$-wave expression is caused by the gain in magnetic energy of the $d$-wave superconducting state by magnetization of the nodal quasiparticles~\cite{vor06c}. The critical value of the Maki parameter where first order transition appears is therefore $\alpha^* = \sqrt{2} H^{orb}_{c2}/H_P \approx 7.0 (H^{orb}_{c2}/B_0) Z^*$, and, since in-plane $Z^* \sim 0.3$ we have $\alpha^*_{d-wave} \approx 3.0$, significantly exceeding the value of unity expected for {\it dirty} isotropic systems. We apply our results to CeCoIn$_5$ to analyse the behavior of the upper critical field along $a$ and $c$ axes. The fit for the $[100]$ direction is shown in Fig.~\ref{fig:hc2fit}. The best fit of the second order transition and the location of $T^*$ in the experimental range is given by $Z^a \sim 0.5$. For $H || c$ the best fit is given by $Z^c \sim 1.15$. From these two fits we obtain the value for $B_0$ which is approximately $30 \: Tesla$ for both fields orientations, which is an indication that our choice of the FS parameters was reasonably good to describe this system. From this value we find the Fermi velocity $v_f \sim 0.6 \cdot 10^6 \: cm/s$, which is sensible for a heavy fermion system and agrees with interpretation of the experimental data~\cite{RMovshovich:2001}, and disagrees with the value $\sim 10^8 \: cm/s$ obtained by authors of Ref.\onlinecite{won04a}. We then extract the values of the effective electron moment, $\mu/\mu_B = Z (2\pi k_B T_c)/(\mu_B B_0)$ and obtain $\mu_{ab}/\mu_B \approx 0.35$ and $\mu_{c}/\mu_B \approx 0.7$, in agreement with Ref.\onlinecite{won04a} where the varying parameters were Fermi velocity and the $g$-factor. We do note that in the current model the orbital pairbreaking is stronger for the nodal direction, Fig.~\ref{fig:PD}(a), hence the Pauli effect is relatively less important, and the range of first order transition is smaller for that orientation, (Fig.\ref{fig:PD}(c)). However, this particular aspect depends sensitively on the in-plane shape of the Fermi surface. We took the Fermi surface to be rotationally symmetric, and the question of how this conclusion is affected by a more realistic Fermi surface shape is left for future studies. \begin{figure}[t] \centerline{\includegraphics[height=5.5cm]{Fig3.eps}} \caption{(Color online) A fit of the experimental $H_{c2} \, // \, [100]$ in CeCoIn$_5$. The data are taken from measurements of magnetization by T.~Tayama \et\cite{TTayama:2002} and specific heat by A.~Bianchi \et\cite{ABianchi:2003}. We find that $Z=0.5$ gives the best fit of second order transition line and the onset of the first order transition $T^*$. } \label{fig:hc2fit} \end{figure} \section{Transport and specific heat anisotropy} We are now in the position to analyse how the Pauli limiting affects the measured anisotropies of the specific heat, $C(\phi_0)$, and longitudinal thermal conductivity, $\kappa_{xx}(\phi_0)$, when magnetic field is rotated in the basal plane with respect to the crystalline axes. The direction of the field, $\phi_0$, is measured from the gap's maximum along the $a$-axis. As is well known, both quantities show oscillations, with either minima or maxima when the field is aligned with the nodal directions~\cite{IVekhter:1999R,AVorontsov:2006,ABVorontsov:2007dos,ABVorontsov:2007kap,YMatsuda:2006,TSakakibara:2007}. The location of the inversion line in the $T$-$H$ plane that separates the regions with minima and maxima indicating the gap nodes is of exceptional experimental relevance as it affects the conclusions about the shape of the gap in a given compound. The results are most clearly presented in the form of a phase diagram in the $T$-$H$ variables that indicates each of the regions. In Fig.~\ref{fig:PD} we show the evolution of such phase diagram for the specific heat anisotropy with increasing Zeeman coupling. We compute the density of states, $N(\vare, T)=-(N_0/2\pi)\langle\Im g(\hvp,\vare)\rangle_{\hvp}$, where inclusion of $T$ signifies that the self-consistently determined gap is temperature-dependent, obtain the entropy, \begin{eqnarray} S(T,\bm H)&=&-\int_{-\infty}^{\infty} d\omega N(\vare, T) \biggl[f(\vare)\ln f(\vare)\nonumber\\&& \qquad\qquad + (1-f(\vare))\ln(1-f(\vare))\biggr]\,, \label{entropy} \end{eqnarray} and numerically differentiate it to find $C/T$. The result for purely orbital coupling to magnetic field, $Z=0.0$ in Fig.\ref{fig:PD}(a) is in agreement with Ref.~\onlinecite{ABVorontsov:2007dos}: In the two shaded regions (near $H_{c2}$ and in low-$T$ low-$H$ corner) the heat capacity attains its minimum when the field is along the nodal directions of the gap. Clearly, since the anisotropy in the heat capacity is field-induced, the amplitude of the anisotropy is very small at low fields for all temperatures, and extremely small near $T_c$. Hence the challenge is to go to sufficiently high fields to see the signal while knowing whether the maximum or a minimum of the anisotropic signal corresponds to the nodal directions. \begin{figure}[t] \centerline{\includegraphics[height=5.5cm]{Fig4.eps}} \caption{(Color online) The anisotropy of heat capacity phase diagram for three Zeeman couplings $Z=0.0$ (a), $Z=0.2$ (b) and $Z=0.4$ (c). Shaded areas correspond to minimum of the heat capacity for $\vH || (node)$. Panels (c1) and (c2) demonstrate how anisotropy of heat capacity, $C(\phi_0)$, varies with temperature and magnetic field for $Z=0.4$. The curves for different $T/T_c$ are shifted vertically for clarity. } \label{fig:PD} \end{figure} As the relative strength of the spin coupling increases, Fig.~\ref{fig:PD}(b) and (c), we note that the ``nodal minimum'' region near $H_{c2}$ grows at the expense of the intermediate field unshaded region. This is concomitant with the disappearance of the anisotropy of the upper critical fields for nodal and antinodal directions. An obvious conjecture is that for $Z=0$ at intermediate temperatures and {\em moderately high fields} it is the anisotropy of the upper critical field that controls the anisotropy of the specific heat: since $H_{c2}^{node}<H_{c2}^{antinode}$, at a fixed external field $H/H_{c2}^{node}>H/H_{c2}^{antinode}$, and the density of states, along with the heat capacity, is higher for the field along the nodal direction. As the Zeeman coupling is increased and the critical fields along the nodal and the antinodal directions become nearly equal, this unshaded region shrinks. In Fig.~\ref{fig:PD}(c) we tuned $Z>Z^*$, and see that the critical fields for the two in-plane directions are almost indistinguishable, and that there is a region of first order transition at low temperatures $T<T^*$, indicated by the arrows. Importantly for our purposes, the low-$T$, low-$H$ shaded region (minima for the field along the nodes) remains mostly unchanged with increasing $Z$, terminating at $T/T_c \sim 0.1$ and $H/H_{c2}^Z \sim 0.4$. This region is still dominated by the nodal quasiparticles and ``semiclassical'' physics, $N(0,H) \sim \sqrt{H/H_{c2}^{orb}}$,~\cite{GVolovik:1993,MineevSamokhin} compared to a linear contributions due to Zeeman shifts (for these $Z$'s) and vortex cores. The insets Figs. \ref{fig:PD}(c1) and \ref{fig:PD}(c2) provide image of the heat capacity anisotropy for $Z=0.4$ coupling for two different fields at locations indicated by circles and squares correspondingly. The heat capacity expression that we use to plot the angular dependence, \be C(T,\vH) = \int_{-\infty}^\infty {d\vare \, \vare^2 \over 4 T^2} {N_\uparrow(\vare,T,\vH) + N_\downarrow(\vare,T,\vH) \over \cosh^2(\vare/2T)}\,, \ee is, strictly speaking, valid only at low temperature when the order parameter is essentially temperature-independent. The difference between this expression and the exact result obtained from the entropy differentiation is small far from the superconducting transition, see Refs.~\onlinecite{AVorontsov:2006,ABVorontsov:2007dos}. \begin{figure}[t] \centerline{\includegraphics[height=4cm]{Fig5.eps}} \caption{(Color online) Anisotropy of the specific heat (left) and heat conductivity (right) for $Z=0.0$ and $Z=0.4$ at $T/T_c=0.3$ and various fields. The field values are shown on the left in terms of $Z$-dependent $H_{c2}$. The curves were shifted vertically for clarity. The absolute values of $C$ and $\kappa$ and their anisotropies presented in Figures below. The range of fields in $H/B_0$ for convenience indicated on the right. } \label{fig:anis} \end{figure} We also compute the anisotropy of the thermal conductivity along the crystalline $x$ direction for unitarity scattering (phase shift $\pi/2$) and the normal state scattering rate $\Gamma/2\pi T_c = 0.007$. The thermal conductivity is calculated on equal footing with the density of states and is given by~\cite{IVekhter:1999,ABVorontsov:2007kap} \begin{eqnarray} \frac{\kappa_{xx}(T,H)}{T} &=& \int\limits^{+\infty}_{-\infty} \; \frac{d\vare}{2 T} \frac{\vare^2}{T^2} \cosh^{-2}\frac{\vare}{2T} \\ \nonumber &\times& \Big< v_{f,x}^2 \, N(T, \bm H; \hat{\vp}, \vare) \; \tau_H(T, H; \hat{\vp},\vare)\Big>_{\hvp} \,, \end{eqnarray} with the effective scattering rate \be \frac{1}{2\tau_H} = - \Im \Sigma^R + \sqrt{\pi}{2 \Lambda \over |\tilde{v}_f^\perp|} \frac{\Im[g^R \, W(2\tilde{\vare}\Lambda/|\tilde{v}_f^\perp|)]} {\Im \, g^R} |\Delta_0 \cY|^2 \,. \label{eq:tauH} \ee Here the label $R$ denotes a retarded function, and we use the angle-resolved density of states. Fig.~\ref{fig:anis} shows characteristic profiles of the heat capacity and heat conductivity when the field is rotated in the basal plane. Note that in the left panel we labeled the fields according to the ratio $H/H_{c2}$: the upper critical field changes between $Z=0$ and $Z=0.4$, so that the effective field range is different. As discussed above, the major difference between two values of the Zeeman splitting is that for higher $Z$ there exists a region of shallow minimum at the node ($45^\circ$ in our case) for high fields. The second panel of Fig.~\ref{fig:anis} shows the behavior of the thermal conductivity for the same fields, now labeled in units of $B_0$. Qualitatively, the peak in the angle-dependent $\kappa$ for $Z=0$ at $45^\circ$ disappears at $Z=0.4$, making the dependence of the thermal conductivity more twofold. \begin{figure}[t] \centerline{\includegraphics[width=0.9\columnwidth]{Fig6.eps}} \caption{(Color online) $C_{0\phi}$ and $C_{4\phi}$ coefficients in heat capacity expansion at two different temperatures, $T=0.05T_c$ and $T=0.3T_c$. Positive four-fold coefficient means that the specific heat has minimum for the field along a node in the gap, while negative $C_{4\phi}$ corresponds to the maximum for the field along a node. } \label{fig:coefC} \end{figure} To quantify these trends we follow the approaches taken in experiment and expand both quantities in harmonics of the angle between the field and the heat current, \bea \frac{C(\phi_0,T,H)/T}{(C/T)_N} &= &C_{0\phi} + C_{4\phi} \cos4\phi_0 \,, \label{eq:coefC} \\ \nonumber \frac{\kappa_{xx}(\phi_0,T,H)/T}{(\kappa_{xx}/T)_N} &=& \kappa_{0\phi} + \kappa_{2\phi} \cos2\phi_0 + \kappa_{4\phi} \cos4\phi_0 \,, \label{eq:coefK} \eea and present the evolution of different coefficients with field and temperature. \begin{figure}[t] \centerline{\includegraphics[width=0.9\columnwidth]{Fig7.eps}} \caption{(Color online) The anisotropic component of the specific heat at temperatures $T=0.05T_c$ (left panel) and $T=0.3T_c$ (right panel) for $Z=0$ and $Z=0.4$. } \label{fig:C4vsHc2} \end{figure} First, focus on the behavior of the specific heat as shown in Fig.~\ref{fig:coefC}. At low fields for $Z=0$ the isotropic part, $C_{0\phi}$, shown in panels (a) and (c), exhibits the approximate $\sqrt{H}$ behavior expected of a nodal superconductor~\cite{GVolovik:1993,CKubert:1998SSC}. With increased Zeeman contribution this component of the specific heat becomes more linear in field~\cite{yan98,IVekhter:2001}. This is, of course, simply because the spin-split quasiparticle spectra produce a density of states at zero energy, $N(0,H) \simeq (\mu H)/\Delta_0$, with the prefactor that, for high $\mu$, can be large enough to dominate the sublinear Volovik term already at low fields. At low temperatures $C_{0\phi}$ acquires a positive curvature at high fields, in agreement with Ref.~\onlinecite{ada05}. The qualitative behavior of the anisotropic term,, $C_{4\phi}$, is similar for the two values of $Z$ at low temperature, panel (b), but is distinctly smaller at higher temperature, shown in panel (d), and even changes sign near $H_{c2}$. The comparison becomes even more clear if the amplitude $C_{4\phi}$ is plotted against the upper critical field for given values of Zeeman splitting, as in Fig.~\ref{fig:C4vsHc2}. At low temperature there is essentially no difference between the purely orbital case and that of a moderately strong Zeeman coupling, and hence in this regime the anisotropy depends only on the shape of the gap and the underlying Fermi surface. In contrast, at $T/T_c=0.3$ for the case of weak Zeeman splitting the coefficient does not change sign, as seen already from Fig.~\ref{fig:PD}(a). In contrast, for $Z=0.4$ as shown in Fig.~\ref{fig:PD}(c), there is a change of sign in the anisotropy of the specific heat, due to the ``isotropization'' of the upper critical field, as discussed above. \begin{figure}[t] \centerline{\includegraphics[width=0.9\columnwidth]{Fig8.eps}} \caption{(Color online) Field dependence of the coefficients $\kappa_0$ and $\kappa_2$ in the expansion of the thermal conductivity, Eq.(\ref{eq:coefK}) at $T=0.05T_c$ (top two panels) and $T=0.3T_c$ (bottom panels) for $Z=0$ and $Z=0.4$.} \label{fig:kappa02} \end{figure} \begin{figure}[b] \centerline{\includegraphics[width=0.9\columnwidth]{Fig9.eps}} \caption{(Color online) The anisotropic fourfold term of the thermal conductivity in Eq.(\ref{eq:coefK}) is shown for $T=0.05T_c$ in panels (a) and (c), and for $T=0.3T_c$ in panels (b) and (d) as function of the field, panels (a) and (b) and the reduced field, $H/H_{c2}$, panels (c) and (d).} \label{fig:kappa4} \end{figure} The same trend is seen in the thermal transport anisotropy. The behavior of the average thermal conductivity, $\kappa_0$ and the twofold term responsible for the difference between heat transport parallel and perpendicular to the vortices, is shown in Fig.~\ref{fig:kappa02}. At low fields the scattering of the quasiparticles on the vortices is determined by the vortex concentration, $n\sim H/\Phi_0$, and therefore, at low temperature, only weakly depends on the Zeeman field. As the temperature increases, however, the differences between the two cases become much more pronounced. In particular, it is worth noting that the two-fold symmetry $\kappa_{2\phi}>0$ ($\kappa_{xx}(\vj_h || \vH) > \kappa_{xx}(\vj_h \perp \vH)$) is much more pronounced in the Pauli-limited case. The fourfold ``nodal'' term, Fig.~\ref{fig:kappa4}, shows the behavior broadly similar to that of the anisotropic component of the specific heat. The low-temperature behavior as a function of the reduced field, $H/H_{c2}$ is similar for the cases of weak and strong Pauli limiting, while the behavior at moderate temperatures is very different starting at moderate fields. Once again, the coefficients for $Z=0$ and $Z=0.4$ have the opposite signs starting at $H\sim 0.5 H_{c2}$. \section{Conclusions} \begin{figure}[t] \centerline{\includegraphics[width=0.9\columnwidth]{Fig10.eps}} \caption{(Color online) The phase diagram for the anisotropy of the specific heat under rotated magnetic field for different strength of Pauli pairbreaking term. Shaded areas correspond to minimum of the heat capacity for $\vH || (node)$. Notice that the shaded region at low $T$ and $H$ is identical in all three panels.} \label{fig:PDsummary} \end{figure} In conclusion, we calculated in a $d$-wave superconductor with quasi-cylindrical FS the $H_{c2}$ and its evolution for transition between orbital and Pauli limits. We find that in this system the critical value of the Maki parameter for paramagnetic limit criterion is $\alpha^*=3$. That values depends on both the symmetry of the order parameter and the shape of the Fermi surface, and hence there is no universal threshold value for that parameter that ensures the first order transition in a given system. We also find that the linear extrapolation of the upper critical field in Pauli-limited superconductors from the vicinity of $T_c$ to low temperatures does not accurately predict the value of the orbital critical field. We showed that moderately large Zeeman contribution does not alter the typical behavior of $C-$ and $\kappa-$anisotropy at low temperatures and fields. This is summarized in the phase diagram in Fig.~\ref{fig:PDsummary}. The region at low $T$ and $H$ where the minima in the fourfold terms in the heat capacity and thermal conductivity occur for the field along the nodes is essentially insensitive to the strength of the Zeeman coupling. The differences, of course, occur at higher temperatures and fields. This finding proves that, even in systems with strong paramagnetic contribution where $H_{c2}$ anisotropy in the $a$-$b$ plane is largely absent and therefore cannot be used to infer the nodal structure~\cite{wei06}, the anisotropy of heat conductivity and heat capacity at moderate to low $H$ and $T$ still can help determine the location of the nodes of the order parameter on the Fermi surface. The previous analysis~\cite{AVorontsov:2006} of the inversion of the anisotropy in CeCoIn$_5$ remains valid when the Zeeman term is accounted for, and unequivocally indicates the $d_{x^2-y^2}$ shape of the order parameter. This also provides further support for the interpretation of Ref.~\onlinecite{KAhn:2009} that confirmed the inversion of the specific heat oscillations. \section*{Acknowledgements} This research was supported in part by the US Department of Energy via Grant No. DE-FG02-08ER46492. We are also grateful for the hospitality of the Aspen Center for Physics, where part of this work was done.
\section{Introduction} The star V505 Sagittarii (HD~187949, HR~7571, HIP~97849, WDS~19531-1436) is known as an eclipsing binary with a variable period. Spectral types of its primary and secondary components are A2\,V and G5\,IV, orbital period is 1\fd183 and visual magnitude in maximum 6\fm5 (Chambliss et al. 1993). In 1985, the V505 Sgr was also resolved using speckle interferometry (McAlister et al. 1987a) and several measurements of this 3rd component were published since that time. Mayer (1997) attempted to join the measured times of minima with visual orbit and determined a distance of the system 102\,pc. The 3rd-body orbit with the period of about 40 years seemed well justified until about the year 2000. An abrupt change in more recent data however excludes this simple model --- it is impossible to fit {\em both\/} light-time effect data and the interferometric trajectory assuming three bodies on stable orbits. We thus test a 4th-body hypothesis: a perturbation by low-mass star (i.e., the 4th body), which is not yet visible in the speckle-interferometry data. Such a fourth body was suspected already by Chochol et al.~(2006) due to conspicuous deviations of minimum times from expectations. While we consider the 4th-body model as the main working hypothesis in this paper, we also discuss other possible effects that can produce minima timings variations. The dataset we have for V505~Sgr is described in Section~\ref{sec:data}. We introduce our dynamical model, numerical method, free/dependent parameters and $\chi^2$ metric in Section~\ref{sec:integrator}. The results of our simulations and conclusions are presented in Sections~\ref{sec:results} and~\ref{sec:conclusions}. \section{Observational data}\label{sec:data} \subsection{Speckle interferometry} The available speckle-interferometry data are summarised in Table~\ref{tab:sky}. Most of them were extracted from the Fourth Catalog of Interferometric Measurements of Binary Stars (Hartkopf et al. 2009), but we also added two speckle measurements from SAO BTA 6\,m telescope by E.~Malogolovets (using a speckle camera and a method described in Balega et al. (2002) and Maksimov et al. (2009)) and one direct-imaging measurement, performed at CFHT by S.~Rucinski (using a method described in Rucinski et al. (2007)). We estimated weight factors $w$ and corresponding uncertainties as $\sigma_{\rm sky} = 0.005\,{\rm arcsec}/w$. The values of $\sigma_{\rm sky}$ vary because of different telescopes and techniques were used. Note these measurement errors sometimes cause that some measured points are seemingly `exchanged' --- the position angle does not revolve monotonically. We are aware of a possible $180^\circ$-ambiguity in the speckle measurements, but V505 Sgr is a lucky case: we have one direct measurement by Hipparcos prior to 2000 perihelion passage, and another direct-imaging datum after 2000. We thus can be sure about the shape of the orbit. \begin{table*} \caption{Speckle interferometry data for V505~Sgr, mainly from the 4th Interferometric Catalogue (Hartkopf et al. 2009). P.A., $d$ denote position angle and angular distance between the central pair (1+2) and the 3rd component. Estimated weight factors $w$ and uncertainties $\sigma_{\rm sky} = 0.005\,{\rm arcsec}/w$ correspond to the sizes of telescopes and techniques, which were used to acquire these measurements (1991.25 and 2005 measurements result from direct imaging).} \label{tab:sky} \centering \begin{tabular}{cllll} \tableline\tableline year & P.A. / deg & $d$ / mas & weight & source \\ \tableline 1985.5150 & 189.6 & 302 & 1 & 3.6 m \cr 1985.8425 & 189.8 & 311 & 1 & 3.8 m \cr 1989.3069 & 181.0 & 261 & 1 & 4.0 m \cr 1990.3445 & 176.9 & 246 & 1 & 4.0 m \cr 1991.2500 & 170 & 234 & 0.6 & Hipparcos \cr 1991.3903 & 173.4 & 234 & 1 & 4.0 m \cr 1991.5575 & 174 & 240 & 0.4 & 2.1 m \cr 1991.5602 & 174 & 260 & 0.4 & 2.1 m \cr 1991.7124 & 173.3 & 226 & 1 & 4.0 m \cr 1992.4497 & 171.7 & 214 & 1 & 4.0 m \cr 1992.6961 & 164 & 190 & 0.4 & 2.1 m \cr 1994.7079 & 159.9 & 192 & 1 & 3.8 m \cr 1995.4398 & 152.5 & 169 & 0.6 & 2.5 m \cr 1995.7675 & 154.2 & 177 & 0.3 & 2.5 m \cr 1996.5320 & 145.8 & 149 & 0.3 & 2.5 m \cr 2003.6365 & 236.3 & 152 & 1 & 3.5 m\cr 2005.7948 & 218 & 183 & 0.6 & direct CFHT \cr 2006.1947 & 215.8 & 182 & 1 & 4.0 m\cr 2007.3306 & 212.4 & 210 & 1 & 3.5 m\cr 2007.4927 & 212.0 & 212 & 1 & 6.0 m\cr 2008.4901 & 207.8 & 231 & 1 & 6.0 m\cr 2009.2662 & 204.2 & 247.5 & 1 & 4.0 m\cr \tableline \end{tabular} \end{table*} \subsection{Minima timings}\label{sec:lite_data} We list recent $O-C$ data for the (1+2) binary in Table~\ref{tab:lite}. Only measurements not presented in Chambliss et al. (1993) are included in the table, but we use all of them of course. An uncertainty of a minimum determination is estimated to $\sigma_{\rm lite} = 1\,{\rm min}$ in most cases, only photographic minima and data from Hipparcos were considered worse. Epoch and $O-C$ were calculated using the ephemeris: \begin{equation} {\rm Pri.Min.} = 2433490.483 + 1\fd1828688 \times E\label{eq:ephem} \end{equation} Note there is a freedom in period and base minimum determination. When we compare these $O-C$ measurements to our simulations we use an optimal ephemeris, different from~(\ref{eq:ephem}). \begin{table*} \caption{Minima timings for the eclipsing binary (1+2) in V505~Sgr. Epoch and $O-C$ were calculated using the ephemeris ${\rm Pri.Min.} = 2433490.483 + 1\fd1828688 \times E$. $\sigma_{\rm lite}$ denotes assumed standard uncertainty of the minimum determination. Only newer minima after Chambliss et al. (1993) are listed. The references are: Rovithis-Livaniou \& Rovithis (1992), M\"uyesseroglu et al. (1996), Ibanoglu et al. (2000), Cook et al. (2005), Chochol et al. (2006), Zasche et al. (2009). The last 4~measurements are new. } \label{tab:lite} \centering \begin{tabular}{rrccl} \tableline\tableline $JD-2400000$ & Epoch & $O-C$ / d & $\sigma_{\rm lite}$ / d & source\\ \tableline 48432.4871 & 12632.0 & $+0.0054$ & 0.0007 & R.-L. \\ 48501.0981 & 12690.0 & $+0.0100$ & 0.0021 & Chochol \\ 48858.3253 & 12992.0 & $+0.0109$ & 0.0007 & M\"uyesseroglu \\ 51000.4948 & 14803.0 & $+0.0049$ & 0.0007 & Ibanoglu \\ 51051.3578 & 14846.0 & $+0.0046$ & 0.0007 & '' \\ 51057.2724 & 14851.0 & $+0.0048$ & 0.0007 & '' \\ 51064.3692 & 14857.0 & $+0.0044$ & 0.0007 & '' \\ 52754.6756 & 16286.0 & $-0.0087$ & 0.0007 & Chochol \\ 52843.3891 & 16361.0 & $-0.0103$ & 0.0007 & '' \\ 53263.3029 & 16716.0 & $-0.0150$ & 0.0007 & '' \\ 53525.8969 & 16938.0 & $-0.0178$ & 0.0007 & Cook \\ 53626.4399 & 17023.0 & $-0.0187$ & 0.0007 & Chochol \\ 54267.5469 & 17565.0 & $-0.0266$ & 0.0007 & Zasche \\ 54267.5472 & 17565.0 & $-0.0263$ & 0.0007 & '' \\ 54648.4260 & 17887.0 & $-0.0313$ & 0.0005 & '' \\ 54655.5233 & 17893.0 & $-0.0312$ & 0.0003 & '' \\ 54658.4817 & 17895.5 & $-0.0299$ & 0.0005 & '' \\ 54706.3869 & 17936.0 & $-0.0309$ & 0.0002 & '' \\ 55027.5302 & 18207.5 & $-0.0365$ & 0.0018 & Uhl\'a\v r \\ 55049.4152 & 18226.0 & $-0.0346$ & 0.0002 & '' \\ 55062.4266 & 18237.0 & $-0.0347$ & 0.0011 & \v Smelcer \\ 55068.3400 & 18242.0 & $-0.0357$ & 0.0011 & Uhl\'a\v r \\ \tableline \end{tabular} \end{table*} \subsection{Radial velocities} We use radial-velocity data from Tomkin~(1992), Tab.~4, who measured sharp spectral lines in the 5580 -- 5610\,\AA\ region and attributed them to the 3rd component. The values of $v_{\rm rad3}$ range from $-13$~to $-9\,{\rm km}/{\rm s}$. The width of the lines corresponds to rotational velocity about $v_{\rm rot3} = (20\pm5)\,{\rm km}/{\rm s}$. The uncertainties of the radial-velocity data $\sigma_{\rm rv} = 2\,{\rm km}\,{\rm s}^{-1}$ were estimated from a scatter of the RV measurements close in time and verified by a practical test: we computed a synthetic spectrum with the same resolution as Tomkin (1992) and fitted the lines in question by a Gaussian function. We also checked for possible blends with nearby faint lines. Wide lines in the V505 Sgr spectrum are attributed to the components of the eclipsing pair (1+2). The binary is tight and in all likelihood rotates synchronously, thus the corresponding rotational Doppler broadening is large ($v_{\rm rot1+2} = (100\pm 10)\,{\rm km}/{\rm s}$). The systemic radial velocity of the (1+2)-body is $v_{\rm rad1+2} = (1.9\pm1.4)\,{\rm km}/{\rm s}$. \section{Numerical integrator and $\chi^2$ metric}\label{sec:integrator} In order to model orbital evolution of the multiple-star system V505~Sgr, namely mutual gravitational interactions of all bodies, we use a Bulirsch-St\"oer $N$-body numerical integrator from the SWIFT package (Levison \& Duncan 1994). Our method is quite general --- we can model classical Keplerian orbits, of course, but also non-Keplerian ones (involving 3-body interactions). We are able to search for both bound (elliptical) and unbound (hyperbolic) trajectories. Free parameters of our model are listed in Table~\ref{tab:free}. Hereinafter, we strictly denote individual bodies by numbers: 1, 2 (Algol-type pair), 3 (resolved third component) and 4 to avoid any confusion. Fixed (assumed) parameters are listed in Table~\ref{tab:dependent}. Masses of the first three components are well constrained by photometry and spectroscopy: $m_1 = (2.20 \pm 0.09)\, M_\odot$, $m_2 = (1.15 \pm 0.05)\, M_\odot$, $m_3 = (1.2 \pm 0.1)\, M_\odot$ (Chambliss et al. 1993, Tomkin 1992). We take $m_3$ as a free parameter, thought, because of larger relative uncertainty. When we test 3-body configurations, we have simply $m_4 = 0$. \begin{table*} \caption{Free parameters of our dynamical 4-body model.} \label{tab:free} \centering \begin{tabular}{rll} \tableline\tableline no. & parameter & brief description \\ \tableline 1. & $d$ & distance of the V505 Sgr barycentre \\ 2. & $m_3$ & mass of the 3rd body \\ 3. & $z_{h3}$ & position, (1+2)-centric, epoch $T_0$ \\ 4. & $v_{xh3}$ & velocities, (1+2)-centric, epoch $T_0$\\ 5. & $v_{yh3}$ & \\ 6. & $v_{zh3}$ & \\ 7. & $m_4$ & mass of the 4th body \\ 8. & $x_{h4}$ & positions, (1+2)-centric, epoch $T_0$ \\ 9. & $y_{h4}$ & \\ 10. & $z_{h4}$ & \\ 11. & $v_{xh4}$ & velocities, (1+2)-centric, epoch $T_0$\\ 12. & $v_{yh4}$ & \\ 13. & $v_{zh4}$ & \\ \tableline \end{tabular} \end{table*} \begin{table*} \caption{Fixed (assumed) parameters of our model.} \label{tab:dependent} \centering \begin{tabular}{rll} \tableline\tableline no. & parameter & brief description \\ \tableline 14. & $m_{1+2} = 3.4\,M_\odot$ & mass of the (1+2) body \\ 16. & $x_{h1+2} = 0\,{\rm AU}$ & positions of the (1+2) body, \\ 17. & $y_{h1+2} = 0$ & (1+2)-centric \\ 18. & $z_{h1+2} = 0$ & \\ 19. & $v_{xh1+2} = 0\,{\rm AU}/{\rm day}$ & velocities \\ 20. & $v_{yh1+2} = 0$ & \\ 21. & $v_{zh1+2} = 0$ & \\ 22. & $x_{h3}$ & positions of the 3rd body, \\ 23. & $y_{h3}$ & (1+2)-centric \\ 24. & $T_0 = 2446282.24375\,{\rm JD}$ & UTC time corresponding \\ & (or $2447607.5185$) & to initial conditions \\ \tableline \end{tabular} \end{table*} First, it is often useful to adopt a simplification: 1st and 2nd body can be regarded as a single (1+2) body in our dynamical model. The central pair (1+2) is so compact ($a = 0.033\,{\rm AU}$) and the distance of other components so large, that it behaves like a single body; its equivalent $J_2$ gravitational moment is negligible. Indeed, at distance $r = 10\,{\rm AU}$ \begin{equation} J_2 \simeq {1\over 2} \left({a\over r}\right)^2 {m_1 m_2\over (m_1+m_2)^2} \simeq 10^{-6}\,. \end{equation} This can be confirmed easily by a direct numerical integration. The difference between trajectories computed for three-body (1, 2, 3) and two-body (1+2, 3) configurations is insignificant and always smaller than observational uncertainties (see Figure~\ref{fit_pozorovani_2BODY_V505_Sgr_arcsec}). \begin{figure*} \centering \includegraphics[height=6.2cm]{figs/fit_pozorovani_2BODY_2009_V505_Sgr_arcsec.eps} \includegraphics[height=6.2cm]{figs/fit_pozorovani_2BODY_V505_Sgr_arcsec_DETAIL.eps} \caption{Comparison of two 3rd-body trajectories, computed for three-body (1, 2, 3) and two-body (1+2, 3) configurations. Left: an overview of the trajectories in a 1-centric frame. Right: a detail of the small part of the trajectory, where the difference is visible. Error bars denote speckle-interferometry observations.} \label{fit_pozorovani_2BODY_V505_Sgr_arcsec} \end{figure*} We also make use of the following two constraints: (i)~initial positions $x_{h3}$, $y_{h3}$ and zero time $T_0$ of the 3rd body correspond to a selected speckle-interferometry datum (e.g., the mean of the first two points, or to the third point); (ii)~3rd body initial velocity components are almost tangent to the observed interferometric trajectory in the $(x,y)$ plane. Initial conditions of the integration are specified in an arbitrary (usually 1+2-centric) frame. We then perform a transformation to a barycentric frame. The numerical integration runs in the barycentric Cartesian frame, where $x, y$~axes correspond to sky plane, $z$~axis is oriented from the observer towards the system. We use AU, AU/day units for positions and velocities. We integrate the system forward for 10,000~days, and backward (i.e., with opposite sign of initial velocities) for 20,000~days in order to cover the observational time span. The time step used is $\Delta t = 10$~days and the precision parameter of the BS integrator is $\epsilon = 10^{-8}$. Finally, we transform the output back to the (1+2)-centric frame and linearly interpolate the output data to the exact times of observations. In order to compare the observations to our model we constructed a~$\chi^2$ metric as follows: \begin{equation} \chi^2 = \chi^2_{\rm sky} + \chi^2_{\rm lite} + \chi^2_{\rm rv}\,,\label{eq:chi2} \end{equation} where \begin{equation} \chi^2_{\rm sky} = \sum_{i=1}^{N_{\rm sky}} {\left(x_{h3}' - x_{h3}[i]\right)^2 + \left(y_{h3}' - y_{h3}[i]\right)^2 \over \sigma_{\rm sky}^2[i]}\,. \end{equation} We denote $x_{h3}'$, $y_{h3}'$ (1+2)-centric coordinates of the 3rd body calculated from our model, which were linearly interpolated to the times $t_{\rm sky}[i]$ of observations $x_{h3}[i]$, $y_{h3}[i]$. Distance~$d$ is used to convert angular coordinates to AU. Secondly, \begin{equation} \chi^2_{\rm lite} = \sum_{i=1}^{N_{\rm lite}} {\left(z_{b1+2}' - z_{b1+2}[i]\right)^2 \over \sigma_{\rm lite}^2[i]}\,, \end{equation} where $z_{b1+2}'$ are barycentric coordinates of the (1+2) body computed from our model and interpolated to the times $t_{\rm lite}[i]$ of observations $z_{b1+2}[i]$. Because of freedom in the period determination and freedom in the selection of initial velocities, we have to detrend the light-time effect data (by two LSM fits of $z_{b1+2}'(t)$ and $z_{b1+2}(t)$). Finally, \begin{equation} \chi^2_{\rm rv} = \sum_{i=1}^{N_{\rm rv}} {\left(v_{zh3}'-v_{zh3}[i]\right)^2 \over \sigma_{\rm rv}^2[i]}\,, \end{equation} where we again interpolate our model to the times $t_{\rm rv}[i]$. Note that in case of a 4-body configuration we will attribute the velocities to the 4th body and change this metric correspondingly (see below). Optionally, we can add an {\em artificial\/} function to $\chi^2$ in order to constrain the mass~$m_4$ within reasonable limits, e.g., \begin{equation} \chi^2_{m_4} = \left[ \left(m_4 - {m_{4\rm min} + m_{4\rm max}\over 2}\right) \cdot {2\over m_{4\rm max}-m_{4\rm min}} \right]^{100}\,,\label{eq:mass_limit} \end{equation} with $m_{4\rm min} = 0.1\,M_\odot$, $m_{4\rm max} = 1.2\,M_\odot$. The upper limit follows from the fact that no other bright star is observed in the vicinity of V505 Sgr. A similar expression can be used to constrain the absolute value of velocity $v_4$ (e.g., to be smaller than the escape velocity from the system, otherwise, we often obtain hyperbolic velocities). Occasionally, we use a different metric instead of~(\ref{eq:chi2}): \begin{equation} \chi^2 = w_{\rm sky} \chi^2_{\rm sky} + w_{\rm lite} \chi^2_{\rm lite} + w_{\rm rv} \chi^2_{\rm rv}\label{eq:chi2_weights} \end{equation} with weights $w_{\rm sky} \ge w_{\rm lite}, w_{\rm rv}$, in order to fit the interferometric trajectory better. There are only five points after the periastron passage, which would otherwise have too low statistical significance compared to a lot of light-time data. What can we expect about the 13-dimensional function $\chi^2(d, m_3, z_{h3}, \dots, v_{zh4})$? It will surely have many local minima, which would be statistically almost equivalent. (One can shoot the 4th body from a slightly different position with a slightly different velocity to get almost the same result.) The problem is degenerate in this sense. Clearly, there are strong correlations, e.g., between the mass $m_4$ and the minimal distance of a close encounter (and consequently initial positions/velocities of the 4th body). Minimisation of the $\chi^2$ function is thus a difficult task. We use a simplex algorithm (Press et al. 1997) to save computational resources and to find local minima. However, it is not our goal to find a global minimum of $\chi^2$, because of the degeneracy and the immense size of the parameter space. We anyway do not expect a deep, statistically significant global minimum. Instead, we will choose a set of starting points for the 4th body and look for a subset of {\em allowed\/} solutions. On the other hand, in case we test a 3-body configuration only, the problem is much simpler: the six-dimensional $\chi^2(d, m_3, z_{h3}, v_{xh3}, v_{yh3}, v_{zh3})$ is well-behaved and we may expect to find a unique solution (and its uncertainty). \section{Results}\label{sec:results} In the following subsections we consider and analyse several hypotheses about the nature of the V505 Sgr system: (i)~there are three bodies only in V505 Sgr; (ii)~the 3rd body directly perturbs the central pair; (iii)~a steady mass transfer causes minima timing variations; (iv)~there is modulation of mass transfer by the 3rd body; (v)~a sudden mass transfer occurred around 2000; (vi)~Appelgate's mechanism is operating; (vii)~a 4th body is present (either on a bound or hyperbolic orbit). \subsection{The 3rd body alone on a Keplerian orbit}\label{sec:keplerian_orbit} At first, let us test a standard "null" hypothesis, i.e., only 3rd body exists ($m_4 = 0$). It is possible to fit speckle data {\em alone\/} ($w_{\rm lite} = w_{\rm rv} = 0$) by an elliptical orbit with a $(29\pm 1)$-yr period, especially, if we assume the first two 1985 measurements are erroneous (offset by 50\,mas, see Figure~\ref{simplex3rd_VZDALENOST2_WO_LITE_V505_Sgr_arcsec}, left). The $\chi^2_{\rm sky} = 50$ for this fit and the respective number of data points is $N_{\rm sky} = 20$. (Thought ideally, $\chi^2$ should be comparable to $N$.) Note the $\chi^2_{\rm sky}$ would be much higher, if we include the 1985 measurements: $\chi^2_{\rm sky} = 210$, $N_{\rm sky} = 22$. It means, if these two measurements are not systematic errors, the 29-yr Keplerian orbit is essentially excluded! The two respective measurements were obtained by two {\em different\/} telescopes during two different nights (see McAlister et al. (1987a) and McAlister et al. (1987b)). We checked measurements of another 34 stars in these publications, observed with the {\em same\/} telescope and during the same night as V505 Sgr, and we have found no indication of a wrong plate scale --- all measurements lie on Keplerian ellipses within usual observational uncertainties (5\,mas). We thus belive the 1985 measurements are {\em not\/} erroneous and they should be included in the $\chi^2$ metric. Without additional (non-positional) data it is not possible to distinguish between different inclinations --- there are equivalent low-$I$ and high-$I$ solutions with almost the same $\chi^2 \simeq 50$. Nevertheless, {\em every\/} inclined orbit of the 3rd body has to cause a corresponding light-time effect, otherwise must be considered wrong! Even a slight $I \gtrsim 2^\circ$ inclination would be easily detectable in the light-time effect data (see Figure~\ref{simplex3rd_VZDALENOST2_WO_LITE_V505_Sgr_arcsec}, middle). A period analysis of the $O-C$ data (with Period04 program) also does not show a prominent 29-yr period. On the other hand, there is a clear signal at $P = 39\,{\rm yr}$, with an amplitude of the peak $A = 0.0092\,{\rm d}$. If we assume the $O-C$ data are indeed caused by a light-time effect, there is a strong disagreement of the 29-yr Keplerian orbit with the light-time effect data (and also with radial velocities), even prior to 2000! If we try to fit the whole orbit and light-time effect data {\em together\/}, we would have $\chi^2_{\rm sky} = 107$ and $N_{\rm sky} = 20$, i.e., such an orbit is excluded with a high significance. There are also clear systematic departures between the observed interferometric data and calculated Keplerian orbit. The only possibility is the inclination of the 3rd-body orbit is almost zero $I \lesssim 2^\circ$, so we do not see any light-time effect at all. The observed $O-C$ variations then must caused by an entirely different phenomenon (see next Sections~\ref{sec:period_change} to \ref{sec:applegate_1992} for a detailed discussion). Nevertheless, there still remains a strong disagreement with the observed high radial velocities $v_{{\rm rad}3} \simeq 10\,{\rm km}/{\rm s}$, because a non-inclined orbit should have $v_{{\rm rad}3} \lesssim 1\,{\rm km}/{\rm s}$. We have no solution for this problem (unless there is a 4th body present in the system, see Sections~\ref{sec:optimalizace_VZDALENOST4} to \ref{sec:simplex_4TH_BODY9}). \begin{figure*} \centering \includegraphics[height=4.7cm]{figs/simplex3rd_VZDALENOST2_WO_LITE_V505_Sgr_arcsec.eps} \includegraphics[height=4.7cm]{figs/simplex3rd_VZDALENOST2_WO_LITE_V505_Sgr_LITE_yr_day.eps} \includegraphics[height=4.7cm]{figs/simplex3rd_VZDALENOST2_WO_LITE_V505_Sgr_RV_yr_kms.eps} \caption{Best-fit solution for the trajectory of the 3rd body, which corresponds to speckle interferometry data (but excluding 1985 measurements). Neither light-time effect nor radial velocities were fitted in this this case. Parameters of the 3rd body are: $m_3 = 1.17\,M_\odot$, $x_{h3} = -0.15\,{\rm AU}$, $y_{h3} = -25.9\,{\rm AU}$, $z_{h3} = -0.63\,{\rm AU}$, $v_{xh3} = 0.0037\,{\rm AU}/{\rm day}$, $v_{yh3} = 0.0017\,{\rm AU}/{\rm day}$, $v_{zh3} = 0.0000\,{\rm AU}/{\rm day}$ for $T_0 = 2447607.5185\,{\rm JD}$. The inclination of the orbit is very low in this case ($I = 1.5^\circ$). The resulting $\chi^2_{\rm sky} = 52$, with the number of data points $N_{\rm sky} = 20$. Note there is a strong disagreement of this Keplerian orbit with both $O-C$ data and radial velocities (total $\chi^2 = 1700$, $N = 90$).} \label{simplex3rd_VZDALENOST2_WO_LITE_V505_Sgr_arcsec} \end{figure*} \subsection{Direct perturbation of the 1+2 orbital period by the 3rd body}\label{sec:period_change} One may ask, if the observed variations in minima timings, which correspond to the changes of the period of the order $|\Delta P| \simeq 10^{-5}\,{\rm d}$, could be caused by a {\em direct\/} gravitational perturbation of the tight central pair (1, 2) by the orbiting 3rd body. In periastron, the minimum distance is of the order $\simeq 10\,{\rm AU}$. In order to test this possibility, we use our dynamical model with three bodies 1, 2 and 3 taken separately. A detection of minute changes of the orbital period requires a smaller time step and higher precision of the BS integrator ($\Delta t = 0.01\,{\rm d}$, $\epsilon = 10^{-12}$). The resulting osculating orbital period changes during one periastron passage are shown in Figure~\ref{Pt}. They are much smaller than $\Delta P \lesssim 10^{-7}\,{\rm d}$. An extremely close encounter (within less then 0.1\,AU, which corresponds to 0.001\,arcsec) would be needed to change the orbital period of the tight Algol system substantially. Moreover, {\em anything\/} directly connected with the 3rd body should conform to the 39~year period of the minima timings and this, according to Section~\ref{sec:keplerian_orbit}, is in conflict with any 29-yr Keplerian orbit of the 3rd body. \begin{figure} \centering \includegraphics[width=7.5cm]{figs/Pt.eps} \caption{Simulated osculating orbital period~$P$ of the central binary (bodies 1 and 2), perturbed by the 3rd body. The periastron passage occurred in 2000 and the corresponding change of period is $\Delta P \lesssim 10^{-7}\,{\rm d}$. The observed values of $|\Delta P| \simeq 10^{-5}\,{\rm d} $ are much larger than in this simulation.} \label{Pt} \end{figure} \subsection{Effects of mass transfer between 1 and 2}\label{sec:mass_transfer} Past photometric and spectroscopic observations confirm the central pair of V505 Sgr is a classical semi-detached Algol system, with a less-massive secondary filling its Roche lobe (Chambliss et al. 1993). In case of a conservative mass transfer, the sum of masses is constant \begin{equation} M_1(t) + M_2(t) = K\,,\label{eq:ZZHM} \end{equation} as well as the orbital angular momentum \begin{equation} A(t) M_1^2(t) M_2^2(t) = C\,,\label{eq:ZZMH} \end{equation} where $A(t)$ denotes the actual separation of the stars. We can substitute current masses and separation $A = 7.1\,R_\odot$ (Chambliss et al. 1993) into these equations, compute constants $K$, $C$ and consequently the dependence $A(M_1)$ (see also Figure~\ref{vzdalenost_V505Sgr}) \begin{equation} A(M_1) = C M_1^{-2} (K-M_1)^{-2}\,.\label{eq:A_M_1} \end{equation} A smooth conservative mass transfer should increase orbital period steadily, since in the V505 Sgr case the mass ratio has been reversed already ($M_1 > M_2$). On contrary, we observe an abrupt {\em decrease\/} of the period $\Delta P = -1.2 \times 10^{-5}\,{\rm d}$ after 2000. We thus conclude a simple mass transfer cannot explain the observer minima timings. \begin{figure} \centering \includegraphics[width=7.5cm]{figs/vzdalenost_V505Sgr.eps} \caption{The dependence of separation~$A$ of the eclipsing binary on the mass~$M_1$ of the 1st body, resulting from the conservation of total mass and orbital angular momentum.} \label{vzdalenost_V505Sgr} \end{figure} \subsection{Modulation of mass transfer between 1 and 2 during the 3rd body encounter}\label{sec:roche} In this section, we test if the 3rd body is capable to change the Roche potential of the central binary (bodies 1 and 2) in a such a way, that the mass transfer rate ${\rm d} M/{\rm d} t$ (and consequently ${\rm d} P/{\rm d} t$) changes by a substantial amount. We add a 3rd-body term to the Roche potential \begin{equation} \Omega(x,y,z) = {1\over r_1} + {q\over r_2} + {1\over 2} (1+q) r_3^2 + {q_{\rm 3rd}\over r_{\rm 3rd}}\,,\label{eq:roche_3rd} \end{equation} where $q = M_2/M_1$ denotes the mass ratio and similarly $q_{\rm 3rd} = M_3/M_1$. We see immediately, that relative change of the potential due to the 3rd body at distance $r_{\rm 3rd} \simeq 10\,{\rm AU}$ is $\delta\Omega/\Omega \simeq 10^{-13}$. We do not find it likely, that such a minuscule perturbation of the potential, and thus the related tidal acceleration, could produce significant effects. Consequently, we cannot explain minima timings variations by the modulation of mass transfer. Finally, similarly as in Section~\ref{sec:period_change}, this effect would be also in conflict with a 29-yr Keplerian orbit of the 3rd body. \subsection{A sudden mass transfer of Biermann \& Hall (1973)}\label{sec:biermann_hall_1973} According to Biermann \& Hall (1973) a sudden mass transfer between the Algol components may result in a {\em temporary\/} decrease of the orbital period, even thought mass is flowing from the lighter component to the more massive. In our case, we would need ${\rm d}M/{\rm d}t$ as high as $\simeq 10^{-6}\,M_\odot/{\rm yr}$ to explain period changes $|{\rm d} P/{\rm d} t| \simeq 10^{-6}\,{\rm d}/{\rm yr}$. Such a mass transfer rate seems to be too large compared to theoretical models (Harmanec 1970), ${\rm d}M/{\rm d}t \gtrsim 10^{-6}\,M_\odot/{\rm yr}$ are reached only during a very short interval of time, before the reversal of mass ratio. Another problem of this scenario is that we observe rather smooth periodic variations of the minima timings before 2000, which do not seem to be entirely compatible with this mechanism, which may be more irregular in time. This phenomenon is also rarely confirmed by independent observations. (It would require a very precise photometry on a long time scale, or a spectroscopic confirmation of circumstellar matter.) Today, this mechanism is not generally accepted as a major cause of minima timing variations among Algol-type systems. \subsection{Applegate (1992) magnetic mechanism}\label{sec:applegate_1992} Applegate (1992) proposed a gravitational quadrupole coupling of orbit and shape variations of a magnetically active subgiant (2nd component) can result in variations of the orbital period and hence minima timings. In this scenario, the observed 39-yr period would correspond to the period of the magnetic dynamo. The 2nd (G5~IV) star rotates quickly (1.2\,d), it has a convective envelope in this evolutionary stage and, presumably, there is a differential rotation and operating dynamo, which can result in a sufficiently strong magnetic field ($10^4\,{\rm G}$), necessary for Applegate's mechanism to work. Period changes of the order $\Delta P/P \simeq 10^{-5}$ should also correspond to changes of the luminosity $\Delta L_2/L_2 = 0.1$, in phase with minima timings. Unfortunately, we are not able to confirm this by our photometry (0.01\,mag precision over tens of years would be required). In principle, this mechanism can explain minima timings variations, but it is not clear, why there is an abrupt change after 2000. An independent confirmation is rare and difficult. One of the possibilities might be a spectroscopic observation of magnetically active lines (Ca~II H and K, or Mg~II). This scenario also does not provide any solution for the observed large radial velocities. \subsection{Distance, mass and the 3rd body orbit (prior to 2000)}\label{sec:optimalizace_VZDALENOST4} Hereinafter, we {\em assume\/} minima timings variations are caused mainly by the light-time effect due to the orbiting 3rd body. Because the orbit of the 3rd body prior the periastron passage in 2000 seems unperturbed, we first determine the optimal distance~$d$ of the system, 3rd-body mass~$m_3$ and orbit ($z_3$, $v_{xh3}$, $v_{yh3}$, $v_{zh3}$). We use only the observational data older than 2000 for this purpose. We compute $\chi^2$ values for the following set of initial conditions (we do not use a simplex here): $d \in (95, 105)\,{\rm pc}$, $\Delta d = 1\,{\rm pc}$, $m_3 \in (1.1, 1.3)\,M_\odot$, $\Delta m_3 = 0.1\,M_\odot$, $z_{h3} \in (2.0, 8.0)\,{\rm AU}$, $\Delta z_{h3} = 1.0\,{\rm AU}$, $v_{xh3} \in (0.0033, 0.0040)\,{\rm AU}/{\rm day}$, $v_{yh3} \in (0.0008, 0.0016)\,{\rm AU}/{\rm day}$, $v_{zh3} \in (-0.0018, 0.0012)$ ${\rm AU}/{\rm day}$, $\Delta v_{xh3} = \Delta v_{yh3} = \Delta v_{zh3} = 0.0001\,{\rm AU}/{\rm day}$. The best-fit solution is displayed in Figure~\ref{optimalizace_VZDALENOST4_V505_Sgr_SKY}. Orbital period of the 3rd body is $P = (39\pm 2)\,{\rm yr}$. The resulting distance $d = (102\pm 5)\,{\rm pc}$. This solution is very similar to that in Mayer (1997). The parallactic distance of V505~Sgr given by Hipparcos ($\pi = (8.40\pm0.57)\,{\rm mas}$, $d = 111\hbox{ to }128\,{\rm pc}$, cf., van Leeuwen 2007) is offset and even the error intervals do not overlap. Note that the radial velocities of the order $-10\,{\rm km}/{\rm s}$ measured by Tomkin (1992) cannot be attributed to the 3rd body, which orbital velocity should be much smaller ($(-2.5\pm0.5)\,{\rm km}/{\rm s}$) according to interferometric and light-time effect data. Consequently, we do not fit the velocities in this case ($w_{\rm rv} = 0$), we are going to attribute them to the 4th body (in the next Section~\ref{sec:optimalizace_4TH_BODY_7D_detail}). Finally, it is important to mention that our solution does {\em not\/} depend on the two ("offset") 1985 speckle measurements at all! We can exclude them completely from our considerations and the result would be the same. Our only assumption was that minima timings variations are caused by the light-time effect and this enforces the orbital period of $P \simeq 39\,{\rm yr}$. (But coincidentally, both 1985 measurements fit perfectly this longer-period orbit.) \begin{figure*} \centering \includegraphics[height=4.9cm]{figs/optimalizace_VZDALENOST4_V505_Sgr_SKY.eps} \includegraphics[height=4.9cm]{figs/optimalizace_VZDALENOST4_V505_Sgr_LITE_yr.eps} \includegraphics[height=4.9cm]{figs/optimalizace_VZDALENOST4_V505_Sgr_RV_yr.eps} \caption{Best-fit solution for the orbit of the 3rd body {\em and\/} the light-time effect before 2000: $d = 102\,{\rm pc}$, $m_3 = 1.2\,M_\odot$, $z_{h3} = 4.0\,{\rm AU}$, $v_{xh3} = 0.0036\,{\rm AU}/{\rm day}$, $v_{yh3} = 0.0013\,{\rm AU}/{\rm day}$, $v_{zh3} = -0.0015\,{\rm AU}/{\rm day}$ for $T_0 = 2446282.24375\,{\rm JD}$. The resulting $\chi^2 = 88$, with the total number of data points $N = 46$. The red lines denote differences between observed and calculated data. Radial velocities were not fitted in this case, they are shown for comparison only.} \label{optimalizace_VZDALENOST4_V505_Sgr_SKY} \end{figure*} \subsection{Encounter with a 4th body (a $\chi^2$ map)}\label{sec:optimalizace_4TH_BODY_7D_detail} We next {\em fix\/} initial conditions of the 3rd body according to the results in Section~\ref{sec:optimalizace_VZDALENOST4} and model a perturbation by a 4th body under different geometries. The free parameters of the model are: $m_4$, $x_{h4}$, $y_{h4}$, $z_{h4}$, $v_{xh4}$, $v_{yh4}$, $v_{zh4}$. We include radial-velocity data, but we assume the spectral lines (and corresponding velocities) belong to the 4th body. We scan the following limited set of initial conditions (over 8~million trials): $m_4 \in (0.5, 0.8)\,M_\odot$, $\Delta m_4 = 0.05\,M_\odot$, $x_{h4} \in (38, 45)\,{\rm AU}$, $\Delta x_{h4} = 1.0\,{\rm AU}$, $y_{h4} \in (37, 40)\,{\rm AU}$, $\Delta y_{h4} = 0.5\,{\rm AU}$, $z_{h4} \in (20, 30)\,{\rm AU}$, $\Delta z_{h4} = 1.0\,{\rm AU}$, $v_{xh4} \in (-0.011, -0.005)\,{\rm AU}/{\rm day}$, $v_{yh4} \in (-0.010, -0.005)\,{\rm AU}/{\rm day}$, $v_{zh4} \in (-0.012, -0.006)$ ${\rm AU}/{\rm day}$, $\Delta v_{xh4} = \Delta v_{yh4} = \Delta v_{zh4} = 0.0005\,{\rm AU}/{\rm day}$. A comparison of the best-fit solution with observational data is displayed in Figure~\ref{optimalizace_4TH_BODY_7D_detail_V505_Sgr_SKY}. We use a modified metric~(\ref{eq:chi2_weights}) with $w_{\rm sky} = 10$, $w_{\rm lite} = w_{\rm rv} = 1$. The respective trajectories of the bodies are shown in Figure~\ref{optimalizace_4TH_BODY_7D_detail_XYZ_baryc}. Note, however, that according to the $\chi^2$~map (Figure~\ref{optimalizace_4TH_BODY_7D_detail_optimalizace_x_y_MIN}) there are many local minima, which cannot be distinguished from a statistical point of view, because the values of $\chi^2$ differ only little ($\chi^2 \in [284, 325]$). The corresponding $\chi^2$ probabilities $Q(\chi^2|N)$, that the observed value of $\chi^2 = 340$ (for a given number of degrees of freedom $N = 105$) is that large by chance even for a correct model, are too low (essentially zero). It may also indicate that real uncertainties might be a bit larger (by a factor of 2) than the values estimated by us. Nevertheless, we will find better solutions using a simplex method (in Section~\ref{sec:simplex_4TH_BODY9}). \begin{figure*} \centering \includegraphics[height=4.8cm]{figs/optimalizace_4TH_BODY_7D_detail_V505_Sgr_SKY.eps} \includegraphics[height=4.8cm]{figs/optimalizace_4TH_BODY_7D_detail_V505_Sgr_LITE_yr.eps} \includegraphics[height=4.8cm]{figs/optimalizace_4TH_BODY_7D_detail_V505_Sgr_RV_yr.eps} \caption{Best-fit solution for the trajectory of the 4th body, which best explains the observed trajectory of the 3rd body, light-time effect and radial velocities: $m_4 = 0.8\,M_\odot$, $x_{h4} = 45.0\,{\rm AU}$, $y_{h4} = 39.5\,{\rm AU}$, $z_{h4} = 28.0\,{\rm AU}$, $v_{xh4} = -0.0105\,{\rm AU}/{\rm day}$, $v_{yh4} = -0.008\,{\rm AU}/{\rm day}$, $v_{zh4} = -0.0075\,{\rm AU}/{\rm day}$. The resulting $\chi^2 = 331$, with the total number of data points $N = 102$. The motion of the 4th body captured in the left panel spans from 1994 to 2003. The squares connected by a straight line indicate the closest encounter between the 3rd and 4th body.} \label{optimalizace_4TH_BODY_7D_detail_V505_Sgr_SKY} \end{figure*} \begin{figure} \centering \includegraphics[height=5.2cm]{figs/optimalizace_4TH_BODY_7D_detail_XYZ_baryc.eps} \caption{Orbits of bodies (1+2), 3, 4 in a barycentric frame and their projection to the plane ($z = -40\,{\rm AU}$).} \label{optimalizace_4TH_BODY_7D_detail_XYZ_baryc} \end{figure} \begin{figure} \centering \includegraphics[width=7.5cm]{figs/optimalizace_4TH_BODY_7D_detail_optimalizace_x_y_MIN.eps} \caption{Minima of the 7-dimensional function $\chi^2(m_4, x_{h4}, y_{h4}, z_{h4}, v_{xh4},$ $ v_{yh4}, v_{zh4})$ at given positions ($x_{h4}, y_{h4}$). The minimum is taken over remaining free parameters $m_4, z_{h4}, v_{xh4}, v_{yh4}, v_{zh4}$. Cross is an overall minimum, black dots represent computed points.} \label{optimalizace_4TH_BODY_7D_detail_optimalizace_x_y_MIN} \end{figure} \subsection{Encounter with a 4th body (different geometry, simplex)}\label{sec:simplex_4TH_BODY9} We selected a different set of initial conditions for the following modelling. They serve as starting points for the simplex algorithm: $m_4 = 0.5\,M_\odot$, $x_{h4} \in (-100, -10)\,{\rm AU}$, $y_{h4} \in (-50.1, -0.1)\,{\rm AU}$, $z_{h4} \in (0, 50)\,{\rm AU}$, $\Delta x_{h4} = \Delta y_{h4} = \Delta z_{h4} = 5.0\,{\rm AU}$, $v_{xh4} \in (0.005, 0.015)\,{\rm AU}/{\rm day}$, $\Delta v_{xh4} = 0.001\,{\rm AU}/{\rm day}$, $v_{yh4} \in (0, 0.01)\,{\rm AU}/{\rm day}$, $\Delta v_{yh4} = 0.002\,{\rm AU}/{\rm day}$, $v_{zh4} \in (-0.007, 0)\,{\rm AU}/{\rm day}$, $\Delta v_{zh4} = 0.001\,{\rm AU}/{\rm day}$. The total number of trials reaches $10^6$. We reject radial velocity constraints ($w_{\rm rv} = 0$), although we can find a lot of allowed solutions with velocities in the correct range ($v_{zh4} \doteq -0.008\,{\rm AU}/{\rm day}$). On the other hand, we use a mass limit according to Eq.~(\ref{eq:mass_limit}). An example of a typical good fit is shown in Figure~\ref{simplex_4TH_BODY9_V505_Sgr_SKY}. We selected one with mass around $m_4 = 0.6\,M_\odot$; the corresponding $\chi^2 = 168$, $N = 73$ and probability $Q(\chi^2|N) \simeq 10^{-9}$, still too low. This solution can be further improved by a 15-dimensional simplex (i.e., with all parameters of the 3rd body free) to reach $\chi^2$ as low as 130 and $Q(\chi^2|N)$ as high as $10^{-5}$. As before, there are many solutions, which are statistically equivalent. We present allowed solutions in Figure~\ref{simplex_4TH_BODY9_simplex_x_chi2_DETAIL_eps} as plots $\chi^2$ versus a free parameter, with each dot representing one local minimum found by simplex. Prominent concentrations of solutions in these plots can be regarded as an indication of more probable solutions. Only minority of trials were successful. Most of them were stopped too early (at high~$\chi^2$) due to numerous local minima. According to the histogram of masses~$m_4$ (Figure~\ref{simplex_4TH_BODY9_m_hist_rozumne_vmax}, left) the values $m_4 < 0.5\,M_\odot$ are less probable and the histogram peaks around $m_4 = 0.9\,M_\odot$. Note the simplex sometimes tends to `drift' to zero or large masses, which leads to artificial peaks at the limits of the allowed interval. The same applies to velocity~$v_4$. Histogram of total energies~$E_4$ of the 4th body (Figure~\ref{simplex_4TH_BODY9_m_hist_rozumne_vmax}, middle) shows a strong preference for hyperbolic orbits ($E_4 > 0$), but elliptic orbits ($E_4 < 0$) also exist (with a 1\,\% probability and slightly larger best $\chi^2 = 199$). The reason for this preference stems from the fact that 3rd body orbit seems almost unperturbed prior to 2000, so one needs rather a higher-velocity encounter of the 4th body from larger initial distance. Typical minimum distances between the 4th and 3rd body during an encounter are around $d_{{\rm min}3} \simeq 6\,{\rm AU}$ and they are even smaller between the 4th body and (1+2) body $d_{{\rm min}1+2} \simeq 1.5\,{\rm AU}$ (Figure~\ref{simplex_4TH_BODY9_m_hist_rozumne_vmax}, right). They are of comparable size and consequently a simple impulse approximation, i.e., an instantaneous change of orbital velocity, cannot be used to link the two elliptic orbits of the 3rd body (before and after the perturbation). There are no good solutions (with $\chi^2 < 300$), which would lead to an escape of the 3rd body. \begin{figure*} \centering \includegraphics[height=4.8cm]{figs/simplex_4TH_BODY9_V505_Sgr_SKY.eps} \includegraphics[height=4.8cm]{figs/simplex_4TH_BODY9_V505_Sgr_LITE_yr.eps} \includegraphics[height=4.8cm]{figs/simplex_4TH_BODY9_V505_Sgr_RV_yr.eps} \caption{A typical solution (out of many) for the trajectory and light-time effect (without radial velocities): $m_4 = 0.576\,M_\odot$, $x_{h4} = 27.912\,{\rm AU}$, $y_{h4} = -84.809\,{\rm AU}$, $z_{h4} = 30.482\,{\rm AU}$, $v_{xh4} = -0.00584\,{\rm AU}/{\rm day}$, $v_{yh4} = 0.01612\,{\rm AU}/{\rm day}$, $v_{zh4} = -0.00620\,{\rm AU}/{\rm day}$. The corresponding $\chi^2 = 168$, with the number of data points $N = 73$. This solution can be further improved by a 15-dimensional simplex to reach $\chi^2$ as low as 130.} \label{simplex_4TH_BODY9_V505_Sgr_SKY} \end{figure*} \begin{figure*} \centering \includegraphics[height=4.8cm]{figs/simplex_4TH_BODY9_simplex_x_chi2_DETAIL_eps.eps1.eps} \includegraphics[height=4.8cm]{figs/simplex_4TH_BODY9_simplex_y_chi2_DETAIL_eps.eps1.eps} \includegraphics[height=4.8cm]{figs/simplex_4TH_BODY9_simplex_z_chi2_DETAIL_eps.eps1.eps} \includegraphics[height=4.8cm]{figs/simplex_4TH_BODY9_simplex_vx_chi2_DETAIL_eps.eps1.eps} \includegraphics[height=4.8cm]{figs/simplex_4TH_BODY9_simplex_vy_chi2_DETAIL_eps.eps1.eps} \includegraphics[height=4.8cm]{figs/simplex_4TH_BODY9_simplex_vz_chi2_DETAIL_eps.eps1.eps} \caption{Distribution of good solutions ($\chi^2 < 300$) in the space of free parameters: $x_{h4}, y_{h4}, z_{h4}, v_{xh4},$ $ v_{yh4}, v_{zh4}$.} \label{simplex_4TH_BODY9_simplex_x_chi2_DETAIL_eps} \end{figure*} \begin{figure*} \centering \includegraphics[height=4.8cm]{figs/simplex_4TH_BODY9_m_hist_rozumne_vmax.eps} \includegraphics[height=4.8cm]{figs/simplex_4TH_BODY9_energy_hist_rozumne.eps} \includegraphics[height=4.8cm]{figs/simplex_4TH_BODY9_DISTANCES_D_MIN12_D_MIN3_hist_rozumne_vmax_ONLY.eps} \caption{Left: Histogram of masses~$m_4$ of all trials and good ones (with $\chi^2 < 300$ and lower absolute value of velocity $|v_4| < 100\,{\rm km}/{\rm s}$). Middle: Histogram of total energies~$E_4$ of the 4th body at the epoch~$T_0$. There is a strong preference for hyperbolic orbits ($E_4 > 0$), but one can find approximately 1\,\% of orbits with negative energies. Right: Histogram of minimum distances $d_{{\rm min}1+2}$, $d_{{\rm min}3}$ of the 4th body from the (1+2) and 3rd body (only good trials are shown). According to these distributions, it is more likely, that the 4th body approached the (1+2) body closer than the 3rd body. Median values are $\bar d_{{\rm min}1+2} = 1.6\,{\rm AU}$ and $\bar d_{{\rm min}3} = 5.9\,{\rm AU}$.} \label{simplex_4TH_BODY9_m_hist_rozumne_vmax} \end{figure*} \subsection{Observational limits of interferometry and CFHT imaging}\label{sec:observational_limits} Postulating an existence of a 4th body, inferred from its gravitational influence on the V505 Sgr system, we should check if this object could have been directly observed in the past. According to A.~Tokovinin (personal communication) the limit of recent interferometric measurements can be approximated by a linear dependence of brightness difference~$\delta y$ in Str\"omgren $y$ magnitudes on angular separation~$d$ of the components: $\Delta y = 4.7\,{\rm mag}$ at $d = 0.15\,{\rm arcsec}$ and $\Delta y = 6.5\,{\rm mag}$ at $d = 1\,{\rm arcsec}$ Adaptive optics at CFHT can reach even fainter. The limit in $K$ band is given by Rucinski et al. (2007), Fig.~7 as a non-linear dependence~$\Delta K(d)$. We can easily select solutions from Section~\ref{sec:simplex_4TH_BODY9}, which fulfil {\em both\/} limits, albeit a lot of them is excluded by the CFHT limit (see Figure~\ref{obs_limit_CFHT}). Note we are not able to predict exact magnitudes or positions of the 4th body, because there are still many solutions possible. Note there is an object in the USNO-A2.0 catalogue, very close to V505 Sgr: 0750-19281506, ${\rm RA}_{\rm J2000} = 298.277034^\circ$, ${\rm DE}_{\rm J2000} = -14.603839^\circ$. This corresponds to an angular separation $2.6$\,arcsec and position angle~$235^\circ$ with respect to V505 Sgr, at the epoch of observation 1951.574. The magnitudes $R = 10.9\,{\rm mag}$ and $B = 11.8\,{\rm mag}$ are marked as uncertain (since the object is located in the area flooded by light of V505 Sgr). This is an interesting coincidence with "our" 4th body, but we doubt the source is real. Moreover, if the brightness of the USNO source is correct within $\pm1\,{\rm mag}$, it should be above the observational limits. \begin{figure*} \centering \includegraphics[height=5.2cm]{figs/obs_limit_interf.eps1.eps} \includegraphics[height=5.2cm]{figs/obs_limit_CFHT.eps1.eps} \caption{Observational limits (black lines) of interferometric measurements (left panel) and the CFHT imaging (right panel). Our solutions from Section~\ref{sec:simplex_4TH_BODY9} are plotted as dots. Many of them are well below the observational limits.} \label{obs_limit_CFHT} \end{figure*} \subsection{Constraints from spectral lines radial-velocity measurements} In previous Section~\ref{sec:optimalizace_4TH_BODY_7D_detail}, we tried to attribute the observed high radial velocities to a hypothetic 4th component. We thus have to ask a question: could the low-mass 4th component be visible in the spectrum? To this end, we used a grid of synthetic spectra based on Kurucz model atmospheres, which was calculated and provided for general use by Dr.~J. Kub\'at (for details of the calculations, cf., e.g., Harmanec et al. 1997). We calculate synthetic spectra for 3 and 4 lights (stars) and compare them with the spectrum observed by Tomkin (1992), Fig.~2. This spectrum was taken at HJD = 2444862.588, close to the primary eclipse of the central binary, which decreases the luminosity of the 1st component and thus weak narrow lines of the 3rd (or 4th) component are more prominent. Modelling of spectra (relative intensities) requires a number of parameters: luminosities, effective temperatures, surface gravity, rotational and radial velocities. Luminosities of the known components (out of eclipse) are: $L_1 = 26\,L_\odot$, $L_2 = 3.8\,L_\odot$, $L_3 = 2.1\,L_\odot$. The amplitude of the lightcurve is $\Delta m = 1.1\,{\rm mag}$ (Chambliss et al. 1993). The effective temperatures are approximately (Popper 1980): $T_{{\rm eff}1} \simeq 9000\,{\rm K}$ (corresponding to A2~V spectral type), $T_{{\rm eff}2} \simeq 6000\,{\rm K}$ (F8~IV to G6--8~IV), $T_{{\rm eff}3} \simeq 6000\,{\rm K}$ (F8~V). We assume the following values of the surface gravitational acceleration: $\log g_1 = \log g_2 = 4.0$ (cgs units), $\log g_3 = 4.5$ (valid for stars close to the main sequence). Rotational velocities of the 1st and 2nd components, a semi-contact binary with an orbital period 1.2~day, are synchronised by tidal lock and are of the order $v_{{\rm rot}1} \simeq v_{{\rm rot}2} \simeq 100\,{\rm km}/{\rm s}$. These are in concert with the observed width of broad spectral lines $\Delta\lambda = 6$\,\AA. For the 3rd component, we assume a lower velocity $v_{{\rm rot}3} = 20\,{\rm km}/{\rm s}$, usual for main-sequence stars. This matches the width of sharp lines. Radial velocities of the 1st and 2nd components are close to zero because of the eclipse proximity ($v_{{\rm rad}1} = v_{{\rm rad}2} \doteq 0$). We assume the following reasonable parameters for the 4th component: $T_4 = 4000\,{\rm K}$ or $5700\,{\rm K}$, $\log g_4 = 4.0$, $v_{{\rm rot}4} = 20\,{\rm km}/{\rm s}$. We construct a $\chi^2$ metric \begin{equation} \chi^2 = \sum_{i=1}^{N_{\rm obs}} {(I_{\rm obs}[i] - I')^2\over\sigma_{\rm obs}[i]^2}\,, \end{equation} where $I_{\rm obs}[i]$ denote observed relative intensities, $\sigma_{\rm obs}[i]$ associated uncertainties and $I'$ is a sum of synthetic intensities weighted by luminosities \begin{equation} I' = {\sum_{j=1}^4 I'(T_{{\rm eff}j}, \log g_j, v_{{\rm rot}j}) \cdot L_j \over \sum_{j=1}^4 L_j} \end{equation} and of course Doppler shifted due to radial velocities ($\lambda' = \lambda_{\rm obs}[i] (1 - v_{{\rm rad}j}/c)$) and interpolated to the required wavelengths $\lambda'$ using Hermite polynomials (Hill 1982). We use a simple eclipse modelling: we decrease $L_1$ according to the Pogson equation to get the observed total magnitude increase~$\Delta m$. Errors $\sigma_{\rm obs}[i]$ were estimated from the scatter in small continua, $\sigma = 0.01$. Artificially small errors $\sigma = 0.003$ were assigned to the measurements in the cores of the narrow lines, in order to match precisely their depths. We constructed a simplex algorithm (Press et al.~1997) with the following free parameters: $L_4$, $v_{{\rm rad}3}$, $v_{{\rm rad}4}$. Other luminosities and radial velocities remain fixed. This simplex is well-behaved and converges to final values almost regardless of starting point. There is no reasonable improvement, if we let all 8~parameters ($L_j$, $v_{{\rm rad}j}$) to be free. The results for two different temperatures are shown in Figure~\ref{3_and_4_lights}. The best fit for $T_4 = 4000\,K$ is $L_4 = 0.22\,L_\odot$, and it is marginally better that the fit with 3~lights only (i.e., with fixed $L_4 = 0$). The luminosity corresponds roughly to the mass $m_4/M_\odot \propto (L_4/L_\odot)^{1/4} = 0.68$, which seems reasonable with respect to the results in Section~\ref{sec:simplex_4TH_BODY9}. Note we used $v_{{\rm rot}3} = 20\,{\rm km}/{\rm s}$ for rotational velocity of the 3rd body. No reasonable solution was found for $v_{{\rm rot}3}$ as high as $100\,{\rm km}/{\rm s}$, which would cause a strong rotational broadening and almost a `disappearance' of spectral lines of the 3rd body. It means, that a low-mass 4th body {\em alone\/} cannot produce deep sharp lines. We thus suspect, there is a blend of lines in the spectrum of observed by Tomkin (1992), which may originate on the 3rd and 4th body, with low and high radial velocities. However, observations with high spectral resolution would be needed to resolve such blending. \begin{figure} \centering \includegraphics[width=7.5cm]{figs/simplex_GRID_T4_4000K_JADRA_CAR_chi2_porovnani_L4_0.eps} \caption{A synthetic spectrum with intensity normalised to continuum for 4 and 3 lights (thick and thin lines) and for the temperature of the 4th body equal to 4000\,K, compared with the observed spectrum (gray line, which thickness corresponds to observational uncertainties), taken from Tomkin (1992). The wavelength range, which we fitted, was 5580\,\AA\ to 5607\,\AA. The best solution for the luminosity of the 4th body is $L_4 = 0.22\,L_\odot$ (with $\chi^2 = 465$, $N = 239$) The fit with 3~lights only is worse, with $\chi^2$ value equal to 542. For $T_4 = 5700\,K$ we would have $L_4 = 0.43\,L_\odot$ and $\chi^2 = 495$.} \label{3_and_4_lights} \end{figure} \subsection{Constraints from the stellar evolution of the eclipsing binary} To assess the long-term evolution of V505 Sgr, we need some information about the age of the system. An {\em upper\/} limit for the age can be estimated easily from masses of stars. The semi-detached central binary (bodies 1 and 2) has a total mass $(3.4\pm0.1)\,M_\odot$. In order to evolve into the current stage, when the 2nd {\em lighter\/} component fills its Roche lobe, the original mass of the 2nd star had to be at least slightly larger than half of the total mass, i.e., $M_2 > 1.7\,M_\odot$. The evolution of radius is shown in Figure~\ref{Rt_M1.7}; we are mainly concerned with the large increase of radius, when the star leaves main sequence. Given the uncertainties of the masses and unknown metallicities, the upper limit for the age is $(2.0\pm0.5)\,{\rm Gyr}$. In order to find a lower limit, we have to check a minimum separation of the components first (cf.~Eq.~\ref{eq:A_M_1} and Figure~\ref{vzdalenost_V505Sgr}). A minimum separation occurs when $M_1 = 0.5 K$, in our case $A_{\rm min} = (5.9\pm0.1)\,R_\odot$. This value is larger than the radius of a $3.4\,M_\odot$ star during the whole evolution on the main sequence. Thus the mass transfer had to start later, in the red-giant phase. The maximum mass of the 2nd star had to be slightly below the total mass, i.e., $M_2 < 3.4\,M_\odot$. According to the $R(t)$ dependence (Figure~\ref{Rt_M1.7}), the red-giant phase starts at the age of $(0.26\pm0.03)\,{\rm Gyr}$, which could be considered as a {\em lower\/} limit for the age of the V505 Sgr system. \begin{figure*} \centering \includegraphics[width=9cm]{figs/Rt_M1.7_M3.4_bw.eps} \caption{Evolution of radii for stars with different masses ($M = 1.6, 1.7, 1.8$ and $3.3, 3.4, 3.5\,M_\odot$, $Z = 0.03$) and metallicities ($M = 1.7$ and $3.4\,M_\odot$, $Z = 0.01, 0.02$). This computation was performed by the program EZ (Paxton 2004). The upper limit for the age of the 2nd component is then 1.5 to 2.5\,Gyr, while the lower limit is between 0.24 and 0.29\,Gyr.} \label{Rt_M1.7} \end{figure*} \section{Conclusions}\label{sec:conclusions} Generally speaking, we are able to explain the observed orbit of the 3rd body together minima timings and radial velocities by a low-mass 4th body, which encounters the observed three bodies with a suitable geometry. There is no unique solution, but rather a set of allowed solutions for the trajectory of the hypothetic 4th body. It is quite difficult to find a solution for both speckle-interferometry and light-time effect data. There are a few systematic discrepancies at the $2\sigma$~level, which cause the likelihood of the hypothesis to be low. Possibly, realistic uncertainties $\sigma_{\rm sky}$ are slightly larger (by a factor 1.5) than the errors estimated by us. Of course, there are other hypotheses, which do not need a 4th body at all (a sudden mass transfer, Applegate's mechanism, etc.), but none of them provides a unified solution for {\em all\/} observational data we have for V505~Sgr. Further observations of the light time-effect during the next decade can significantly constrain the model. A new determination of the systemic velocity of V505 Sgr may confirm, that the change in the $O-C$ data after 2000 resulted from an external perturbation. (Tomkin's (1992) value was $(1.9\pm 1.4)\,{\rm km}/{\rm s}$.) Spectroscopic measurements of the indicative sharp lines would be also very helpful to resolve the problem with radial velocities mentioned in the text. If we indeed observe the V505 Sgr system by chance during the phase of a close encounter with a 4th star, we can imagine several scenarios for its origin: \begin{enumerate} \item A random passing star approaching V505 Sgr on a hyperbolic orbit. The problem of this scenario is a very low number density of stars. If we take the value $n_\star \simeq 0.073\,{\rm pc}^{-3}$ from the solar vicinity (Fern\'andez 2005), the mean velocity with respect to other stars of the order $v_{\rm rel} \simeq 10\,{\rm km}/{\rm s}$ and the required minimum distance of the order $d \simeq 10^2\,{\rm AU}$, we end up with a mean time between two encounters $\tau \simeq {1 / (n_\star v_{\rm rel} d^2)} \simeq 10^{12}\,{\rm yr}$, thus an extremely unlikely event. \item A loosely bound star on a highly eccentric orbit, with the same age as other three components of V505 Sgr. Unfortunately, there is a large number of revolutions and encounters ($10^2$ to $10^5$) over the estimated age of V505 Sgr and the system practically cannot remain stable over this time scale (Valtonen \& Mikkola 1991). \item A more tightly bound star on a lower-eccentricity orbit, which experienced some sort of a late instability, induced by long-term evolution due to galactic tides, distant passing stars, which shifted an initially stable configuration into an unstable state, e.g., driven by mutual gravitational resonances between components. The problem in this case is that tightly bound orbits of the 4th body are very rare in our simulations, thus seem improbable. \end{enumerate} None of the scenarios is satisfactory. Nevertheless, we find the 4th-body hypothesis the only one which is able to explain all available observations. Clearly, more observations and theoretical effort is needed to better understand the V505 Sagittarii system. \begin{acknowledgements} This research has made use of the WDS Catalog maintained at the U.S. Naval Observatory and Aladin and Simbad services. We thank W.I.~Hartkopf for sending us recent positional data of double stars, S.~Rucinski and E.~Malogolovets for providing us with their measurements of the visual component of V505~Sgr (by CFHT and SAO BTA telescopes) and L.~\v Smelcer for his photometry of V505~Sgr. We acknowledge the use of a grid of synthetic spectra prepared by J.~Kub\'at and we thank P.~Harmanec for some advice on the construction of synthetic spectra for multiple systems. This work was supported by the Czech Science Foundation (grant no.~P209/10/0715) and the Research Programme MSM0021620860 of the Czech Ministry of Education. TP~acknowledges support from the EU in the FP6 MC ToK project MTKD-CT-2006-042514. \end{acknowledgements}
\section{Introduction} The formation and evolution of galaxies is still an open question. This fundamental problem can be tackled by looking at the ages and abundances of the member stars of nearby galaxies (i.e. galactic archeology). Information about a galaxy's formation and subsequent evolution can be inferred by looking at the spatial distribution and stellar population parameters (e.g.,ages and metallicities) of its constituent stars as they are observed today. This can be done accurately via studying the composition of individual member stars. Unfortunately, except for a handful of very nearby galaxies, it is not possible to resolve individual stars even with the most powerful telescopes available today. One must therefore find means of extracting information about the star formation and enrichment history from the integrated light of stellar populations. This is easiest done for globular clusters (GCs), which are well approximated as single stellar populations. Using integrated light spectra of GCs together with either empirical relationships or single stellar population (SSP) models, one can estimate stellar population parameters such as ages and metallicities. Moreover, GCs are believed to have formed early in the history of the universe because they are typically measured to be old (e.g.,Forbes \& Forte 2001; Kuntschner et al. 2002; Brodie et al. 2005; Strader et al. 2005; Cenarro et al. 2007, hereafter C07; Proctor et al. 2008). The study of GC systems provides information about the early star formation history of their host galaxies and can be used to constrain galaxy formation scenarios. For example, GC systems typically show a bimodal color distribution. Although there has been some debate on its origin (Yoon, Yi \& Lee 2006), the currently favored interpretation is that the bimodality in color reflects a metallicity bimodality because of the predominately old ages of GCs (Brodie \& Strader 2006). Therefore, the blue and red subpopulations correspond to a metal-poor and metal-rich subpopulation, respectively. Any galaxy formation scenario must provide at least two different star formation epochs or mechanisms to account for the metallicity bimodality of the host's GC system (but see Muratov \& Gnedin 2010). Early-type galaxies typically have large numbers of GCs and are thus an excellent target for GC studies. For this reason, the present study concentrates on the elliptical galaxy NGC 1407, a brightest group galaxy (BGG) dominating the Eridanus A group (Brough et al. 2006). It harbours a rich GC system (Perrett et al. 1997; Harris et al. 2006; Forbes et al. 2006; Spitler et al. 2010, in preparation). As with other BGGs and brightest cluster galaxies (e.g.,Harris 2009a), NGC 1407's blue GC subpopulation shows a trend of color with luminosity such that brighter GCs have redder colors on average. This `blue tilt' has been observed in several, usually massive, galaxies of different morphological types and is usually interpreted as a mass-metallicity relationship (see Strader et al. 2006; Harris et al. 2006; Mieske et al. 2006; Spitler et al. 2006; Strader \& Smith 2008; Bailin \& Harris 2009; Forbes et al. 2010; Blakeslee et al. 2010). Although this interpretation has been questioned by some (Kundu 2008; Waters et al. 2009), its reality has been confirmed by Harris (2009b) and Peng et al. (2009). In order to assess the formation history, enrichment history, confirm the origin of the color bimodality and the blue tilt in GC systems, statistically significant samples of spectroscopically determined metallicities of GCs are required. Spectroscopic data of large GC samples are rare and typically time consuming to acquire. However, using the DEIMOS multi-object spectrograph on Keck it is possible to obtain over 100 integrated light spectra of GCs simultaneously. Using DEIMOS, we have obtained over 100 spectra of kinematically confirmed GCs around NGC 1407 suitable for stellar population analysis. The sensitivity of DEIMOS is good at red wavelengths near the region of the Ca\thinspace\textsc{ii} triplet (hereafter CaT, 8498, 8542, and 8662 \AA). The CaT is known to correlate with metallicity for integrated light spectroscopy of Galactic GCs (Bica \& Alloin 1987; Armandroff \& Zinn 1988). The CaT has also been studied as a potential metallicity indicator for integrated light of \emph{galaxies} in the past with varying degrees of success. The word `puzzle' has been put forward with regards to its `unexpected' behavior with metallicity. For example, the CaT was found to be lower than predicted by theory in giant elliptical galaxies (Saglia et al. 2002) and higher than predicted in dwarf elliptical galaxies (Michielsen et al. 2003). Possible resolution of the puzzle has been obtained by comparing CaT strengths to metallicities determined using optical spectra in dwarf ellipticals rather than metallicities derived from narrow-band photometry (Michielsen et al. 2007). Nevertheless, Cenarro, Cardiel \& Gorgas (2008) and Foster et al. (2009) were able to successfully use the CaT to probe the metallicity gradients of M32 and a sample of massive to intermediate mass elliptical galaxies, respectively. Therefore, while the behavior of the CaT with respect to metallicity has been studied for the integrated light spectra of galaxies, which are composite stellar populations, with varying degrees of success, it is worth investigating whether it can be used straightforwardly as a metallicity indicator for the integrated light of simple stellar populations such as extragalactic GCs. A description of the data used in this work is presented in Section \ref{sec:data}. In Section \ref{sec:analysis} we analyze the observational and theoretical behavior of the CaT with metallicity and present our choice of CaT index definition. Results can be found in Section \ref{sec:results}. Finally, a discussion and our conclusions are given in Sections \ref{sec:discussion} and \ref{sec:conclusions}, respectively. \section{Data}\label{sec:data} \subsection{Photometry} The photometric data consist of imaging of the central region (3.4 x 3.4 arcmin) from the Advanced Camera for Surveys (ACS) mounted on the \emph{Hubble Space Telescope} (HST) with both the $F435W$ ($B$) and $F814W$ ($I$) bands. The ACS dataset has been independently analyzed and published by both Forbes et al. (2006) and Harris et al. (2006). Here this is supplemented by Subaru/Suprime-Cam images covering a wider field of view (34 x 27 arcmin) in the SDSS $g$, $r$, and $i$ filters (see Spitler et al. 2010, in preparation). Globular cluster candidates in the central region imaged by both Suprime-Cam and ACS thus have photometry in the $B$, $I$, $g$, $r$, and $i$ filters while those outside the ACS field only have $g$-, $r$-, and $i$-band photometry. In all cases, the reddening corrections were performed according to the DIRBE dust maps (Schlegel, Finkbeiner \& Davis 1998). Both Forbes et al. (2006) and Harris et al. (2006) find a relationship between color and luminosity for the blue GC subpopulation (blue tilt) in their ACS imaging of NGC 1407. Kundu (2008) and Waters et al. (2009) have argued that the blue tilt in M87 (also a massive elliptical galaxy) could be a photometric bias caused by the resolved sizes of the brightest metal-poor GCs in the HST/ACS images as opposed to an intrinsic astrophysical phenomenon. However, these claims have recently been refuted by both Harris (2009b) and Peng et al. (2009). The Subaru/Suprime-Cam color magnitude diagram (CMD) for all photometrically selected GC candidates brighter than $i=23.0$ around NGC 1407 is shown in Figure \ref{fig:cmd}. This apparent magnitude limit corresponds to an absolute magnitude of $M_i=-8.9$ when assuming the average redshift independent distance modulus $(m-M)=31.9$ given in NED\footnote{NASA/IPAC Extragalactic Database (NED) is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration.}. The CMD reveals a bimodal GC distribution as found by Forbes et al. (2006) and Harris et al. (2006). The photometric blue tilt is apparent and there is no equivalent red tilt. We also show our spectroscopic subsample and compute the running average for both the red and blue subpopulations. The distribution of our selected spectroscopic subsample in the CMD is representative of the underlying GC color distribution for $i\le22.0$ (or $M_i\le-9.9$). In order to obtain $(B-I)_{0}$ colors and \emph{combine} both datasets (i.e. HST/ACS and Subaru/Suprime-Cam) we have used the candidates in the common central region to derive the following empirical conversion between $(g-i)_{0}$ and $(B-I)_{0}$ colors: \begin{equation} (B-I)_{0}=(1.40\pm0.05)(g-i)_{0}+(0.49\pm0.04). \end{equation} This conversion is useful when comparison our results to single stellar population models that do not always provide SDSS colors (see Section \ref{sec:models}). Figure \ref{fig:color_conv} shows the color conversion as well as the residuals with projected galactocentric radius. The residuals are slightly higher at small projected galactocentric radii due to crowding and the more uncertain Subaru photometry near the center where NGC 1407's surface brightness is high. However, as we move to larger projected radii the error in the converted $(B-I)_{0}$ is reduced to only a few hundredth of a mag (see the lower panel of Figure \ref{fig:color_conv}). In Figure \ref{fig:BIhist} we show both the color histogram of our combined data and the spectroscopic subsample, whose color distribution is representative of that of the combined data. In order to directly compare with previous literature, we apply the heteroscedastic (unequal widths) KMM test to all candidates within galactocentric radii $\le2$ arcmin (comparable to the ACS field). This yields $(B-I)_{0}=1.62$ and 2.07 for the mean color of the blue and red GC subpopulations, respectively. These values are in good agreement with the previous comparable analyses based on ACS imaging of Harris et al. (2006) and Forbes et al. (2006) who found blue peaks at 1.63 and 1.61, and red peaks at 2.07 and 2.06, respectively. However, because of radial color gradients similar to those found by Harris (2009b) in M87, the larger Subaru field of view has peaks at $(B-I)_0=1.57$ and 2.00 with widths of 0.15 and 0.16 for the blue and red GC subpopulations around NGC 1407, respectively. Using the SSP models of Vazdekis et al. (2003, hereafter V03) the difference in the average $(B-I)_0$ colors for the two subpopulations translates into a metallicity difference of $\Delta$[Fe/H]$\approx1.0$ for a fixed old age ($\sim13$ Gyr) or an age difference of $> 8$ Gyr assuming a fixed moderate metallicity of [Fe/H]$=-0.38$ for all GCs. The latter can be ruled out because an age difference as large as 8 Gyr would have already been detected spectroscopically (see C07). We refer the reader to Spitler et al. (2010, in preparation) for further details on the photometry. \begin{figure} \epsscale{1.19} \plotone{f1.eps} \caption{Colour magnitude diagram for NGC 1407 GCs brighter than $i=23.0$ from Subaru/Suprime-Cam data. Grey points show the position of all photometrically selected GC candidates (including a small number of potential NGC 1400 GCs). Red and blue points show the spectroscopically confirmed GCs with CaT measurements. A running average (solid blue and red lines) of our sample has been overplotted for both GC subpopulations indicating a blue tilt but no red tilt. The color division is $(g-i)_{0} = 0.93$ following Romanowsky et al. (2008). The top x-axis shows the color conversion which we derive in Figure \ref{fig:color_conv} for comparison. }\label{fig:cmd} \end{figure} \begin{figure} \epsscale{1.19} \plotone{f2.eps} \caption{Empirical color conversion from $(g-i)_{0}$ to $(B-I)_{0}$. The top panel shows the least squares linear fit to the data. One outlier due to crowding was removed from the fit and is not shown. Differences between converted and actual $(B-I)_{0}$ colors with galactocentric radius are shown underneath. These residuals decrease with radius down to a few hundredth of a mag for radii $\gtrsim1.0$ arcmin.}\label{fig:color_conv} \end{figure} \begin{figure} \epsscale{1.19} \plotone{f3.eps} \caption{Colour histogram of our combined (Subaru/Suprime-Cam and HST/ACS) photometric data for all GC candidates with $i\le23.0$ (black) and our spectroscopic subsample scaled by a factor of 10 for clarity (grey). Overlaid are the Gaussian profiles estimated by the KMM test (assuming heteroscedasticity) with mean blue and red peaks at $(B-I)_{0}=1.57$ and 2.00, respectively, for the whole photometric sample.}\label{fig:BIhist} \end{figure} \subsection{Spectroscopy} \subsubsection{Acquisition} The Keck/DEIMOS spectroscopic data of our photometrically selected GC candidates was obtained during two distinct observing runs. First, three masks were observed on the nights of 2006 November 19-21. The 1200 l mm$^{-1}$ grating with 7500 \AA\space central wavelength and 1 arcsec slit width was used. Details of the first observing run can be found in Romanowsky et al. (2009). The second observing run occurred on the nights of the 2007 November 12-14 under good seeing conditions (typically $\sim 0.7$ arcsec). Three masks were observed with similar instrumental setup as that of Romanowsky et al. (2009). The 1200 l mm$^{-1}$ grating centred on 7800 \AA\space was used. The setup for both observing runs allows coverage of the wavelength range $\sim7550-8900$ \AA\space at redder wavelengths with a resolution of $\sim1.5$\AA\space around the CaT features. Using this setup, we also occasionally cover H$_{\rm \alpha}$ (6563\AA) at bluer wavelengths for a small fraction of our objects. Four 30 minute exposures were taken on each mask, yielding a total exposure time of 2 hours. \subsubsection{Data reduction}\label{sec:reduction} The reduction of the 2006 data is described in Romanowsky et al. (2009). As with the 2006 data, the \textsc{idl} spec2d data reduction pipeline written for the DEEP2 galaxy survey was used to reduce the 2007 DEIMOS data. The spectra were then cross-correlated with the solar spectrum to extract radial velocities using the \textsc{fxcor} routine in \textsc{iraf}\footnote{\textsc{iraf} is distributed by the National Optical Astronomy Observatories, which are operated by the Association of Universities for Research in Astronomy, Inc., under the cooperative agreement with the National Science Foundation.}. Using these velocities, it was possible to clearly distinguish between GCs belonging to NGC 1407, background unresolved galaxies and foreground stars. One mask included some NGC 1400 GCs, but those were easily identified because of the significantly different recession velocity of NGC 1400 (558 km s$^{-1}$, NED) and its GC system from that of NGC 1407 (1779 km s$^{-1}$, NED) as seen from Figure \ref{fig:rvhist} (see also Romanowsky et al. 2009). The velocity selection yielded 113 new confirmed GCs around NGC 1407. These data were supplemented by those of Romanowsky et al. (2009) to give a total of 274 distinct confirmed GCs. The radial velocities of the 7 objects in common between both the 2006 and 2007 data show very good agreement with an rms scatter of only 11 km s$^{-1}$. Because the NIR region of the spectrum is strongly affected by skylines we find that residual skylines in our raw spectra leftover from the sky subtraction influence our index measurements. In order to counteract this, the spectra are template fitted with the \textsc{pPXF} code of Capellari \& Emsellem (2004) using 13 stellar templates that were observed during the 2007 November run with the same instrumental setup as the GC data. The templates include 11 giant and 2 dwarf stars spanning spectral types from F to early M and a wide range in CaT depth. Known problematic regions of the spectrum with strong skylines were excluded during the fitting procedure (see Proctor et al. 2009; Foster et al. 2009). The \textsc{pPXF} routine redshifts, broadens, and chooses the weighted combination of the templates that minimizes the residuals between the raw spectrum and the fit. The routine is unable to fit the noisiest spectra as well as the incomplete spectra affected by vignetting due to their position near the edges of the DEIMOS mask. These $\sim 20$ spectra are not used further for index measurements. The final \textsc{pPXF} fitted spectra are the ``best'' skyline residual free description of our spectra assuming only minimal template mismatches. This technique is similar, and more accurate in principle, than fitting purely Gaussian profiles to the CaT features (Battaglia et al. 2008). Figure \ref{fig:templates} shows examples of template fitted spectra for one of the brightest and one of the faintest GCs in our final sample. The residuals are mostly uniform but showing some features that are mostly associated with the position of known skylines, indicating that template mismatch is not significant. Next, the best fit spectra are continuum normalized interactively using the \textsc{iraf} \textsc{continuum} routine with a spline3 function of order typically $\sim 4$ and a higher and (stricter) lower sigma clipping to ensure that spectral features are not fitted. This sets the continuum to unity. In what follows, we will refer to these \textsc{pPXF} and continuum normalized spectra as ``fitted spectra'' in order to distinguish them from the ``raw spectra'' output from the DEIMOS pipeline. Finally, the fitted spectra with a raw average number of counts less than 80 were removed from the sample. This corresponds to an apparent/absolute $i$-band magnitude and signal-to-noise ratio (S/N) cut of approximately 22.0/--9.9 mag and 9 per \AA, respectively (see Appendix \ref{Appendix:errors}). The final sample contains 144 GCs associated with NGC 1407. The index values are measured on both the fitted and raw spectra. Determination of the error on the index values is discussed in Appendix \ref{Appendix:errors}. \section{Analysis}\label{sec:analysis} In this section, we first review the observational evidence for the sensitivity of the CaT to metallicity in the integrated light spectra of GCs. We then describe and motivate our choice of CaT index definition. We apply our index definition to model spectra from V03 and Bruzual \& Charlot (2003, hereafter BC03) in order to better understand theoretically the behavior of the CaT with metallicity. These will form the observational and theoretical bases on which our results are obtained and discussed. \begin{figure} \epsscale{1.19} \plotone{f4.eps} \caption{Histogram showing the recession velocity of all confirmed GCs around NGC 1400 and NGC 1407. The systemic velocities of NGC 1400 and NGC 1407 are labelled to show that their respective GC systems are easily delineated in velocity space with a gap at $\approx1000$ km s$^{-1}$.}\label{fig:rvhist} \end{figure} \begin{figure} \epsscale{1.19} \plotone{f5.eps} \caption{Panels (a) and (b) show the raw spectra of a sample bright ($i=20.20$) and a faint ($i=21.98$) GC, respectively. Overlaid in red are the fitted spectra. Residuals, mostly caused by skylines (yellow highlight) and noise, are shown underneath for each fit. Regions shaded in grey show the CaT index positions.}\label{fig:templates} \end{figure} \subsection{Revisiting Armandroff \& Zinn (1988)}\label{sec:AZ88} The CaT was first recognised as a potential metallicity indicator in the works of Bica \& Alloin (1987) and Armandroff \& Zinn (1988, hereafter AZ88). In particular, the latter obtained integrated spectra of Galactic GCs to measure the CaT index and averaged literature metallicity measurements. After removing the data points with uncertain metallicity measurements they used linear regression to fit a straight line to the remaining 7 data points (i.e. 47 Tuc, NGC 362, NGC 1851, NGC 5927, NGC 6093, M15, and M2). They obtained the linear relationship: \begin{equation} \mathrm{[Fe/H]}=-4.146 + 0.561 \times \mathrm{CaT_{AZ88}}, \label{eq:AZ88} \end{equation} with rms scatter of only 0.12 dex. This relationship is shown in Figure \ref{fig:AZ88}. The metallicities are the average between the quoted values in the literature at the time and their CaT$_\mathrm{AZ88}$ derived values. There exists more recent metallicity measurements for several GCs and it is worthwhile to verify that the then observed tight correlation still holds with these newer data. In Figure \ref{fig:AZ88} the new measurements as compiled in the 2003 updated version of the Catalogue of Parameters for Milky Way Globular Clusters (Harris 1996, hereafter H96) are also shown. In what follows, we use [Fe/H] to denote metallicities although the metallicities given in both AZ88 and H96 (and hence herein) are based on the Zinn \& West (1884) metallicity scale, which is not a strict Fe scale (e.g.,Carreta \& Gratton 1997; Rutledge, Hesser \& Stetson 1997). Immediately striking in Figure \ref{fig:AZ88} are the four circled data points. These are four of the eight GCs for which AZ88 were the first to give a metallicity estimate. The GCs are HP 1 (Ortolani, Bica \& Barbuy 1997a), Terzan 1 (Ortolani et al. 1999a), Terzan 4 (Ortolani, Barbuy \& Bica 1997b; Origlia \& Rich 2004), and Terzan 9 (Ortolani et al. 1999b). All four are bulge GCs, therefore estimates of their metallicity are plagued by extinction and contamination by foreground bulge stars. In fact, as explained in the above respective references, it is possible that the CaT$_\mathrm{AZ88}$ measurements were contaminated by foreground metal-rich bulge stars. More particularly, as explained in Barbuy et al. (2006), the metallicity of HP1 is a matter of debate with metallicity estimates differing by as much as 1.2 dex. In any case, these four uncertain data points were not originally used by AZ88, leaving the relatively tight relationship unaltered. A linear fit to the updated data for the original 7 GCs selected by AZ88 (filled stars in Figure \ref{fig:AZ88}) yields results consistent with that of AZ88 within $1\sigma$, indicating that more recent metallicity measurements have left the relationship unchanged. However, the most metal-rich GC (NGC 5927) used for this analysis has a metallicity of $-0.4$ dex and it is unclear whether the relationship remains linear beyond this point. Nevertheless, this confirms that it is still justified to obtain metallicities for Galactic GCs from the CaT$_\mathrm{AZ88}$ index measurements using the AZ88 relationship at least for [Fe/H]$\lesssim-0.4$ and possibly beyond. \begin{figure} \epsscale{1.19} \plotone{f6.eps} \caption{An updated reproduction of Figure 5 of AZ88 showing the tight correlation between CaT$_\mathrm{AZ88}$ and [Fe/H]. CaT$_\mathrm{AZ88}$ index values are taken directly from AZ88. Hollow squares are the metallicity measurements quoted by AZ88. Stars represent updated [Fe/H] values from H96 (2003 update) with filled symbols showing the 7 points chosen by AZ88 to determine their relationship (solid line). Dotted lines are drawn between the corresponding measurements for a given GC to guide the eye. The four circled outliers are discussed in the text. Recent metallicity measurements have left the original relationship found by AZ88 essentially unchanged for [Fe/H]$\lesssim-0.4$.}\label{fig:AZ88} \end{figure} \subsection{Choice of index definition}\label{sec:IndexDef} Several definitions for the CaT index have been used in the literature. Among the definitions that apply to integrated spectroscopy are those of AZ88; Diaz, Terlevich \& Terlevich (1989, hereafter DTT89); and Cenarro et al. (2001, hereafter C01). For a review covering these index definitions and how they compare see C01. As mentioned above, AZ88 used integrated spectra of GCs to measure the CaT index. Their definition (CaT$_\mathrm{AZ88}$) therefore should be suitable for GC integrated light spectra such as the present dataset. The CaT index definition of DTT89 (CaT$_\mathrm{DTT89}$) is slightly broader, thus more suitable for the study of galaxy integrated light for which velocity dispersion broadens the CaT features. One drawback of this definition is that it uses the same two continuum passbands for all three CaT lines and is thus strongly affected by changes in the shape of the continuum. Because hot stars (B, A, and F types) have pronounced Paschen lines, three of which overlap with the CaT, one has to worry about contamination of the CaT by Paschen lines from such stars. Therefore, when studying the behavior of the CaT in a wide range of stellar temperatures the C01 definition (CaT$^{*}_\mathrm{C01}$) may be preferable because it is designed to correct for the Paschen line contribution. Provided that the GCs in NGC 1407 are comparable to their Galactic counterparts, one can use the empirical correlation obtained by AZ88 to obtain metallicities. The CaT$_\mathrm{AZ88}$ index central passbands ($Ca1 =$ [8490.0-8506.0\space\AA$]$, $Ca2 =$ [8532.0-8552.0\space\AA], $Ca3 =$ [8653.0-8671.0\space\AA]) were thus adopted for this work. As the typical velocity dispersion of GCs is small, the relatively narrow definition of AZ88 is appropriate. Also, contamination by Paschen lines should not be a concern for integrated spectroscopy of GCs because they are strong only in the theoretical spectra of young stellar populations ($\lesssim 2$ Gyrs, but see Appendix \ref{sec:SmallLines}). Therefore, the passbands of the index definition of AZ88 were selected in order to use their empirical conversion to metallicity. In Figure \ref{fig:indices} we identify some of the features present in the CaT region of the spectrum. Our adopted definition of the CaT index is shown. Because the fitted spectra are continuum normalized, the continuum was set to unity to compute the fitted indices and the continuum passbands of AZ88 were not needed. The CaT indices computed using this method will be referred to simply as CaT in what follows. The continuum normalization has introduced some systematic differences between our index values and those of AZ88. Indeed, like flux calibration, continuum normalization can cause systematic deviations that are difficult to quantify precisely without access to the original AZ88 data (see C01 for a discussion of this). In order to quantify the systematics we measure both the CaT$_\mathrm{AZ88}$ (i.e. measured on the raw spectra) and CaT (i.e. after fitting) indices on the SSP models of Vazdekis et al. (2003) for ages of 8 Gyrs and older. Figure \ref{fig:compare} shows the comparison, which has an offset but a very small scatter. The same is true for our GC spectra, albeit with larger scatter due mostly to the effects of skyline residuals on the measured CaT$_\mathrm{AZ88}$. The equation of the best-fit line to the V03 SSP models points is: \begin{equation} \mathrm{CaT_{AZ88}}=(0.783\pm0.025)\times \mathrm{CaT}+(0.96\pm0.17), \label{eq:convCaT} \end{equation} with $r^2=0.97$ implying that 97 \% of the V03 data are accounted for by the derived relationship. Combining with Equation \ref{eq:AZ88} above yields the following conversion between CaT and [Fe/H]: \begin{equation} \mathrm{[Fe/H]_{CaT}}=-3.641+0.438\times \mathrm{CaT}.\label{eq:convtoFe} \end{equation} We will use Equation \ref{eq:convtoFe} to convert our CaT measurements into metallicities. Another potential source of systematics between our measured indices and those of AZ88 is the difference in resolution (corresponding to $\sigma=51$ km s$^{-1}$ in our case). Fortunately, as shown in Figure 5b of C01, the relative measurement error induced by such a difference in resolution of $\sigma=51$ km s$^{-1}$ for the CaT$_\mathrm{AZ88}$ index is insignificant. \begin{figure} \epsscale{1.19} \plotone{f7.eps} \caption{Example of a fitted spectrum with the main spectral features highlighted. Identified spectral features include the CaT definition (Pa13, Pa15, and Pa16 lines overlap with the Ca3, Ca2, and Ca1, respectively), and the C01 Paschen `triplet' lines (Pa17, Pa14, and Pa12). Three other metallicity sensitive lines are also highlighted (see Appendix \ref{sec:SmallLines}).}\label{fig:indices} \end{figure} \begin{figure} \epsscale{1.19} \plotone{f8.eps} \caption{An assessment of the systematics introduced by our index measurement method. The x-axis shows the CaT index value measured on the fitted spectra while the y-axis shows the same value measured using the AZ88 index definition on the raw spectra. Black and gray points show values measured using old ($>8$ Gyr) SSP model spectra from V03 and our GC spectra with S/N$\ge15$, respectively. The solid line shows the best fit line (Equation \ref{eq:convCaT}) through the model points that is used to correct our index measurements.}\label{fig:compare} \end{figure} \subsection{Single stellar population models}\label{sec:models} SSP models can provide theoretical insight for understanding the behavior of the CaT index. Indeed, GCs are believed to be well approximated as SSPs. One should thus be able to directly compare GC properties with those of SSP models in the literature. Unfortunately, as will be demonstrated below, different SSP model sets make discrepant predictions about the behavior of the CaT index with metallicity. An advantage of using SSP models, particularly those providing spectral energy distributions (SEDs), is that it is possible to perform template fits and continuum normalization in order to measure the CaT index so as to self-consistently directly compare the observed measurements with the model predictions. The present analysis is compared to the V03 and the BC03 SSP models, which both supply SEDs of sufficient resolution. Moreover, the V03 and BC03 models include photometry in the Johnson-Cousins system and a wide range of metallicities and ages. We use the V03 SSP models with a Kroupa (2001) initial mass function (IMF) with metallicities between $-1.68\le$[Fe/H]$\le0.02$. For the BC03 models, we use the SSPs with a Chabrier (see Chabrier 2003) IMF (similar to the Kroupa IMF) in the metallicity range $-1.7\le$[Fe/H]$\le0.00$ so as to directly compare with the V03 models. Models with ages of 5, 9, and 13 Gyrs are selected in both V03 and BC03. Figure \ref{fig:models} shows the predictions of the V03 and BC03 models for CaT versus $(B-I)_{0}$ and metallicity. When no other independent metallicity measurement is available, colors are used as a diagnostic for the sensitivity of the CaT to metallicity. The $(B-I)_{0}$ colors of GCs are known to correlate with metallicity for Galactic and extragalactic GCs (see for example Barmby et al. 2000, hereafter B00; Spitler, Forbes \& Beasley 2008). However, as shown by Smith \& Strader (2007), this conversion is affected by the selected Galactic GC sample from which it is derived due to observational errors and/or interstellar reddening. From Figure \ref{fig:models}, we see that there is broad agreement between the V03 and BC03 SSPs in the overall range of the CaT values and its qualitative sensitivity to metallicity, though they differ greatly in the details. Indeed, the two different sets of models assign widely different metallicities to the same CaT value. A feature that is found in the V03 models is the `loss of sensitivity' or `saturation' of the CaT features to metallicity as the metallicity increases beyond [Fe/H]$\sim$[Fe/H]$_{\mathrm{CaT}}\sim-0.5$ dex for the $\sim13$ Gyrs model. This qualitative behavior is in disagreement with that predicted by the BC03 models for which the CaT sensitivity to metallicity appears to increase for higher values of [Fe/H]. The effect of age on the measured CaT index is minimal for old ages (e.g.,DTT89; V03; Carrera et al. 2007). Indeed, as can be seen on Figure \ref{fig:models} both the V03 and BC03 models show very small variations of the CaT for ages $\gtrsim5$ Gyrs. The empirical calibration of AZ88 (Equation \ref{eq:AZ88}), which is based on Galactic GCs, appears to lie below and above the V03 and BC03 models, respectively, in Figure \ref{fig:models}. We note however that the slope of the V03 models at low metallicities ([Fe/H]$\lesssim-0.5$) is in good agreement with Equation \ref{eq:convtoFe}. This was already shown in Figure 14 of V03. In summary, we find that the BC03 and V03 models make discrepant predictions with respect to the behavior of the CaT as a function of metallicity. For this reason, we choose to use the empirically derived conversion of AZ88 in this work. \begin{figure} \epsscale{1.19} \plotone{f9.eps} \caption{Predictions from V03 and BC03 SSP models. The dotted, dashed, and solid lines correspond to 5, 9, and 13 Gyrs models. From left to right, the red hollow circles on the V03 model lines correspond to [Fe/H]$= -1.68, -1.28, -0.68, -0.38, 0.00, 0.02$ and the blue hollow diamonds on the BC03 model lines correspond to [Fe/H]$= -1.7, -0.7, -0.4, 0.0$. The top x-axis of the left panels shows the [Fe/H]-values for the conversion from $(B-I)_{0}$ by B00 based on Galactic GCs. Our conversion between CaT and $[FeH]$ for [Fe/H]$\le-0.4$ (Equation \ref{eq:convtoFe}) is shown as a dashed gray line for comparison.}\label{fig:models} \end{figure} \section{Results: The GC system of NGC 1407}\label{sec:results} NGC 1407 is a brightest group galaxy (BGG) showing clear GC color bimodality with a division between the blue (metal-poor) and red (metal-rich) GCs occurring around $(B-I)_{0}=1.84$ or $(g-i)_{0}=0.93$ (see Harris et al. 2006; Forbes et al. 2006; Romanowsky et al. 2009). C07 obtained Keck/LRIS spectra of the brightest GC candidates. They derived metallicities and ages for 19 confirmed GCs and 1 ultra compact dwarf (UCD) using the Lick/IDS system (Gorgas et al. 1993; Worthey et al. 1994) and the method of Proctor, Forbes \& Beasley (2004). They found the majority to be old, with 3 being either young (i.e. $\sim 4$ Gyrs) or old GCs with blue horizontal branches (hereafter young/BHB GCs). Figure \ref{fig:C07data} shows that the B00 relationship between color and metallicity (derived for Galactic GCs) is also consistent with the NGC 1407 GC data. Noticeable in Figure \ref{fig:C07data} is the position of the 3 young/BHB GCs, which agrees with the 5 Gyrs V03 model line. However, as mentioned in C07, their position could also be explained by the presence of a blue horizontal branch (BHB) in these clusters. The results presented in Figure \ref{fig:AZ88}, \ref{fig:compare} and \ref{fig:C07data} suggest that, under the assumption that the GCs in NGC 1407 are not significantly different from those of the Milky Way (i.e. similar stellar content), the CaT index values should scale linearly with $(B-I)_{0}$ colors (at least for metallicities below [Fe/H]$\lesssim-0.5$ according to the V03 models). Figure \ref{fig:BICaT} shows the relationship between the $(B-I)_{0}$ colors and CaT for our sample of GCs. The CaT values for the Milky Way GCs were converted from the CaT$_\mathrm{AZ88}$ quoted in AZ88. The agreement between the NGC 1407 dataset and the V03 models is good with the data scattering about the 13 Gyr track. While a general trend between color and CaT is observed, there are several interesting features present. Indeed, it appears as though the CaT index flattens out or `saturates' for metallicities higher than about [Fe/H]$\approx-0.5$ (on the right y-axis) as predicted by V03. This behavior is in contradiction with the prediction of increasing metallicity sensitivity made by BC03 and with an extrapolation of the AZ88 relation to higher metallicities. We therefore decided to consider the V03 models more closely than the BC03 models as they yield better agreement. The bulk of the GCs around NGC 1407 (and the V03 models) apparently have either higher CaT index values, or bluer colors, than the Galactic GCs as shown in Figure \ref{fig:BICaT}. This apparent offset between the Galactic GCs and the NGC 1407 GCs could be related to the lack of flux calibration in both this work and AZ88, which makes the two studies difficult to compare directly. Moreover, the small number and lack of redder Galactic GCs complicates this comparison. Also, the Galactic GCs are plagued by large photometric uncertainties due to Galactic extinction. Indeed, Figure 14 of V03 has already shown that the behavior of the CaT with metallicity in Galactic GCs is in reasonable agreement with their models. Because of the uncertainties involved, the relative position of the Galactic GCs along both axes of Figure \ref{fig:BICaT} is uncertain and we refrain from drawing strong conclusions from it. The GC metallicity distribution obtained by applying our CaT to [Fe/H]$_{\mathrm{CaT}}$ transformation (Equation \ref{eq:convtoFe}) is shown in Figure \ref{fig:tilts}. There are several unexpected features present in the inferred metallicity distribution. First, the blue tilt observed in the CMD (Figure \ref{fig:cmd}) appears to still be present, however there is also a hint of an inverse red tilt. The spread in metallicity of the blue GCs seems to be wider than that of the red GC subpopulation in contrast to their spreads in color that are similar for both subpopulations. Moreover, using the CaT as a metallicity indicator seems to have merged the red and blue bright peaks even though they are separated in color by more than $\Delta (g-i)_{0} \approx 0.2$ or $\Delta (B-I)_{0} \approx 0.25$ mag. This color separation corresponds to an expected metallicity difference of over 0.8 dex for old ages (V03). Figure \ref{fig:brightGC} shows the mean raw spectra of the brightest ($i<20.5$) red and blue GCs. They have similar CaT line strengths and hence inferred metallicity. The same exercise was repeated using the median value of the brightest spectra with, and without fitting, to ensure that the average was not biased by one outlier spectrum, skylines or our template fitting procedure. Each time there was no obvious difference between the strength of the CaT features in the bright red and blue GC spectra. On the other hand, the averaging of the brightest red GCs has increased the signal-to-noise ratio sufficiently to detect the Mg\thinspace{\sc{i}} line at 8807\AA\space while this feature is barely present in the mean spectrum of the brightest blue GCs. We measure the generic Mg\thinspace{\sc{i}} index defined in Cenarro et al. (2009) for both spectra and find that the brightest red GCs have Mg\thinspace{\sc{i}}$=0.77\pm$0.13\AA\space while the brightest blue GCs have Mg\thinspace{\sc{i}}$=0.15\pm$0.15\AA. The Mg\thinspace{\sc{i}} line is mostly sensitive to both temperature and [Mg/H] which correlates with [Fe/H] (see Cenarro et al. 2009). This suggests that the bright blue GCs are either more typical of a hotter (earlier-type) or a less metal enriched stellar population than the red ones. The latter is more likely since hot stars contribute very little at near-infrared wavelengths. At fainter magnitudes a difference in CaT is seen. The KMM test (Ashman, Bird \& Zepf 1994) performed on the CaT inferred metallicities for the \emph{whole sample} shows that a bimodal distribution is preferred over a unimodal one for this dataset at the 96 \% confidence level. The output mean metallicities are [Fe/H]$_{\mathrm{CaT}}=-1.20$ and [Fe/H]$_{\mathrm{CaT}}=-0.61$ for the blue and red subpopulations, respectively. This can arguably be seen in the non-symmetric shape of the [Fe/H] or CaT values histogram (see Figure \ref{fig:histogram}). We are however careful drawing any strong conclusions from these results as: 1) the shape of the distribution seen in Figure \ref{fig:histogram} is considerably different from that seen in the color histogram (Figure \ref{fig:BIhist}), and 2) the numbers are low. Moreover, the confidence level drops below significance (87 \% only) when the KMM test is performed on the CaT$_{\rm AZ88}$ index measured on the raw data possibly due to the larger measurement errors. Nevertheless, with these caveats in mind, we compare with Forbes et al. (2006) who found an average of [Fe/H]$ = -1.45$ and [Fe/H]$ = -0.19$ for the metal-poor and -rich GC subpopulations based on $(B-I)_{0}$ photometry for NGC 1407 GCs. The inconsistently lower mean metallicity for the red subpopulation than that found in the analysis by Forbes et al. (2006) could be explained by: 1) possible radial gradients (see Section \ref{sec:data}), and/or 2) the prediction of V03 of the saturation of the CaT feature around [Fe/H]$\sim-0.5$. However, it appears that the blue GC subpopulation has systematically higher measured CaT than expected. This cannot be attributed to either radial trends or saturation effects. It is puzzling that the absolute position of the blue GCs is shifted towards higher metallicities causing the two subpopulations' metallicity distributions to be similar for the most luminous GCs. \begin{figure} \epsscale{1.19} \plotone{f10.eps} \caption{Relationship between [Fe/H] and color for a sample of NGC 1407 GCs based on blue spectroscopic data from C07. The hollow circles highlight the 3 GCs identified in C07 as potentially young or harbouring BHBs while the triangles show common GCs between this work and that of C07. The solid black line shows the B00 relationship while the solid and dotted red lines represent the 13 and 5 Gyrs V03 model predictions, respectively.}\label{fig:C07data} \end{figure} \begin{figure} \epsscale{1.19} \plotone{f11.eps} \caption{The distribution of NGC 1407 GCs in the $(B-I)_{0}$-CaT plane. Solid and dotted red lines are the V03 models for 13 and 5 Gyr, respectively. Young models of age 1.6 Gyrs are also plotted as a solid grey line. Grey stars are Galactic GCs ($E(B-V) \le 0.3$) with colors taken from H96 and CaT values from AZ88 (after applying the inverse of Equation \ref{eq:convCaT}). Squares show our data with relative sizes proportional to the the signal-to-noise in the raw spectrum. Hollow symbols are discussed in Appendix \ref{sec:SmallLines}. The top x-axis shows the B00 color-metallicity relationship based on Galactic GCs and the right y-axis shows the metallicity derived from CaT using Equation \ref{eq:convtoFe}. Typical error bars shown.}\label{fig:BICaT} \end{figure} \begin{figure} \epsscale{1.19} \plotone{f12.eps} \caption{The inferred metallicity distribution for our sample of NGC 1407 GCs. A running average (solid blue and red lines) have been overplotted for the blue and red subpopulations (colors are as inferred from the photometry). Typical error bars are shown (see Appendix \ref{Appendix:errors}).}\label{fig:tilts} \end{figure} \begin{figure} \epsscale{1.19} \plotone{f13.eps} \caption{Averaged and continuum normalized raw spectra of the brightest ($i\le20.5$) blue and red GCs in NGC 1407. The red spectrum has been shifted by 200 km s$^{-1}$ from a zero redshift for clarity. Even though their average $(g-i)_{0}$ colors are separated by over 0.2 mag, the CaT line strengths of the two mean raw spectra are nearly identical suggesting similar metallicities. This is in contrast with the MgI feature that is deeper in the red GCs as expected for higher metallicity.}\label{fig:brightGC} \end{figure} \begin{figure} \epsscale{1.19} \plotone{f14.eps} \caption{A histogram showing the distribution of the CaT (lower axis) and [Fe/H]$_{\mathrm{CaT}}$ (upper axis) in our sample. The KMM test shows that the distribution of our data is best fitted with two Gaussians (overlaid) at the 96 \% confidence level. The inferred mean metallicities are [Fe/H]$_{\mathrm{CaT}}=-1.20$ and [Fe/H]$_{\mathrm{CaT}}=-0.61$ for the blue and red subpopulations, respectively.}\label{fig:histogram} \end{figure} \section{Discussion}\label{sec:discussion} Based on the results presented in Figures \ref{fig:AZ88}, \ref{fig:compare}, and \ref{fig:C07data}, the CaT index should scale linearly with metallicity and $(B-I)_{0}$ color at least for [Fe/H]$\le-0.4$. On the other hand, single stellar population models disagree about the sensitivity and behavior of the CaT features with respect to metallicity. The empirical calibration of AZ88 was adopted throughout this work to derive metallicities from the CaT index values under the assumption that the GCs around NGC 1407 are intrinsically similar to those around the Galaxy. As demonstrated in Figure \ref{fig:brightGC}, the bright blue and red GCs in NGC 1407 have the same CaT index values. This would suggest that they have the same average metallicity even though they differ in mean $(g-i)_0$ color by over 0.2 mag. The metallicity inferred from the AZ88 relationship for both the brightest blue and red GCs is [Fe/H]$_{\mathrm{CaT}}\sim-0.8$ dex. Moreover, even though the GC system is clearly bimodal in color and in CaT, which is measured on the fitted spectra, it is not bimodal in the CaT$_{\rm AZ88}$ index distribution measured on the raw spectra. This is possibly due to increased measurement errors of the indices measured on the raw spectra. For this reason, we do not consider that our CaT data are sufficient to confidently confirm (i.e. with more than 95 \% confidence) that the GC system around NGC 1407 is bimodal in CaT inferred metallicity contrary to what is expected from the colors. However, in Appendix \ref{sec:SmallLines} we show how the distribution of the sum of the equivalent width of three weak features present in the \emph{fitted spectra only} exhibit clear bimodality. The disagreement between the metallicities inferred from the colors and those inferred from the CaT index values points to a fundamental difference between the GCs around NGC 1407 and the Milky Way either with respect to the behavior of their colors or the CaT index with metallicity. However, it is possible that the data have some systematic biases that could cause this behavior. Such a systematic effect would need to affect the blue GC subpopulation much more than the red one. One possible source of systematics is the sky subtraction. If the sky was oversubtracted for the blue GCs only, this would yield higher CaT measurements. This was investigated and there is no obvious offset between the continuum levels of the blue and red subpopulations. The behavior of the CaT with decreasing signal-to-noise was tested by adding Poisson noise to V03 SSP models SEDs of various metallicities. While lower signal-to-noise spectra inevitably yield to larger errors on the measured CaT index values, no systematic scattering to higher or lower values, or with metallicity or color was found (see Appendix \ref{Appendix:errors}). Assuming the data are reliable we now explore several possible explanations. One possible explanation to the similar metallicities inferred for the bright blue and red GCs is that the CaT features saturate at lower metallicities than predicted by the SSP models (i.e. at [Fe/H]$_{\mathrm{CaT}}\sim-0.8$ instead of [Fe/H]$_{\mathrm{CaT}}\sim-0.5$). If we make the reasonable assumption that our composite near-infrared spectra are not significantly influenced by hot early-type stars, the presence of a stronger Mg\thinspace{\sc{i}} line in the mean spectrum of the bright red GCs than in that of the bright blue GCs does indeed suggest that the CaT could be saturated while the Mg\thinspace{\sc{i}} is still sensitive to metallicity. However, the data presented in Figure \ref{fig:BICaT} suggests a saturation metallicity that is in agreement with the V03 models. Nevertheless, early saturation of the CaT could happen if the GCs around NGC 1407 were Calcium enhanced with respect to their Galactic counterparts (i.e. [Ca/Fe] is larger in NGC 1407 GCs). Indeed, the data presented in Battaglia et al. (2008) for the CaT in individual red giant branch stars suggests that a different [Ca/Fe] ratio than is present in the calibration GCs may influence the measured CaT values and thus the inferred metallicities by up to about 0.2 dex error for [Fe/H]. A saturation of the CaT features at lower metallicities could also be the result of a more bottom-heavy IMF for the GCs around NGC 1407 with respect to those around the Galaxy. Because the CaT index decreases with increasing surface gravity (DTT89; Cenarro et al. 2002), dwarf stars have lower CaT values. A stellar population with a bottom-heavy IMF would contain a larger proportion of dwarf stars at old ages and thus should saturate at lower CaT index compared to a more top-heavy IMF. For example, models with a Salpeter (1955) IMF saturate at higher CaT values than models with a Kroupa (2001) IMF in the V03 SSP models. The Mg\thinspace{\sc{i}} index is only minimally influenced by gravity (Cenarro et al. 2009), so it is unlikely to be significantly influenced by a different IMF and should still be a good tracer of metallicity. The relative spread in CaT inferred metallicity of the blue and red subpopulations are at odds with what is expected from their respective spread in color. If proven correct, this could be due to the non-linearity of the conversion between colors and metallicity (e.g.,Yoon et al. 2006; Peng et al. 2006). In some galaxies, a non-linear conversion between color and metallicity could cause a unimodal metallicity distribution to appear bimodal in color (Cantiello \& Blakeslee 2007; Blakeslee et al. 2010). Therefore, if we take our inferred metallicities at face value, the color and metallicity distributions could be reconciled by invoking this effect. Alternatively, the difference in the relative spreads of the respective subpopulations between the color and CaT distributions could be a result of systematic biases. Indeed, the higher sensitivity of the CaT to metallicity for metal-poor (blue) GCs, which have metallicities lower than the saturation limit predicted by V03, allows for a wider range of inferred metallicities. However, the saturation limit is reached at metallicities corresponding to the metal-rich (red) GCs. This in turn reduces the range of allowed CaT values and thus the inferred range in metallicities for the red GC subpopulation. It is possible, although an extreme case, that the majority of our blue spectra are contaminated by Paschen line absorption. If this were true it could play a definite role in explaining our inferred metallicity distribution. There are several spectral features of the Paschen series of hydrogen in the spectral region of the CaT. Their individual depths vary slightly with the deepest lines at redder wavelengths. Because three of the features of the Paschen series of Hydrogen overlap with the three CaT features, even a small level of contamination or order $\approx 0.3$\AA\space per overlapping Paschen line would be sufficient to significantly alter the measured CaT index by $\approx0.9$\AA. Paschen lines are predominant in stellar spectra of B, A, and F spectral types. These usually massive hot stars are short lived and thus old stellar populations such as GCs are not expected to show significant Paschen absorption features from such stars. Indeed, Paschen lines only become significant in the V03 SSP models for ages $\lesssim2.0$ Gyrs and metallicities [Fe/H]$\lesssim-0.68$. However, keeping in mind that models are uncertain at young ages (V03), we cannot rule out the possibility these GCs could be young. Non-overlapping Paschen lines around the CaT (i.e. Pa12, Pa14 and Pa17) with equivalent widths of order $\approx 0.3$\AA\space would not be detectable in our modest signal-to-noise spectra. However, in Appendix \ref{sec:SmallLines} we discuss how Paschen lines are sometimes visible in the \emph{fitted spectra only}. Another alternative is a population of GCs with hot blue stars such as BHB or blue stragglers stars (Cenarro et al. 2008) at intermediate metallicities. This is consistent with the favored conclusion of C07 regarding the young/BHB GCs in NGC 1407 based on the diagnostic of Schiavon et al. (2004). Moreover, there is some evidence from UV studies that the GC systems of massive elliptical galaxies such as NGC 1407 could harbour a significant population of GCs with extreme hot horizontal branches (see Sohn et al. 2006; Mieske et al. 2008). We examined the archival \emph{Galaxy Evolution Explorer} (GALEX) UV images of NGC 1407 and found that they were not deep enough for its GC system to be detected. If an additional contribution from Paschen line to the CaT index caused by hot blue stars is indeed present, then NGC 1407 could harbour a population of \emph{intermediate metallicity} GCs with BHBs or extreme HBs as speculated from UV studies of other galaxies or a population of GCs with a large proportion of blue straggler stars. With this in mind, it is worth reconsidering Figure \ref{fig:C07data} and the results of C07. The $(r-i)_{0}$ colors should be less affected by, although not immune to, the presence of hot blue stars. This is because hot blue stars contribute less of the integrated light at redder wavelengths (e.g.,Smith \& Strader 2007; Spitler, Forbes \& Beasley 2008; Cantiello \& Blakeslee 2007). We therefore compare the position of the 3 young/BHB GCs identified by C07 in both color spaces. The position of the 3 young/BHB GCs are consistent with the bulk of the other GCs with measured old ages using $(r-i)_0$ colors (right hand panel of Figure \ref{fig:C07data}). In summary, the unexpected distribution of CaT values and inferred metallicities for NGC 1407 GCs could be explained by 1) early saturation of the CaT features in NGC 1407's GCs, 2) a population of GCs with hot blue stars in NGC 1407, and/or 3) the non-linear conversion between color and metallicity. With the current dataset and based on the current generation of SSP models at near-infrared wavelengths, we cannot positively determine which (combination) of these possible explanations is the correct one. Unfortunately, this casts serious doubts on the CaT inferred metallicity distribution shown in Figure \ref{fig:tilts}, the presence of a spectroscopic blue tilt, and the potential of the NIR CaT feature as a metallicity indicator in the integrated light spectra of extragalactic GCs. Until these issues are understood, metallicities inferred from the CaT for integrated light spectroscopy of extragalactic GCs will remain uncertain. \section{Summary and Future work}\label{sec:conclusions} The empirical relationship found by Armandroff \& Zinn (1988) between the CaT$_\mathrm{AZ88}$ index values and metallicities in Galactic GCs was revisited in the light of more recent metallicity measurements. No noticeable difference was found and based on the literature we conclude that the CaT can in principle be used to determine metallicities in integrated spectroscopy of extragalactic GCs. This assumes that the Galactic GCs are not intrinsically different from extragalactic GCs. We also compare the predictions for the behavior of the CaT with metallicity for the V03 and BC03 SSP models. We find that different SSP model sets assign widely different absolute metallicities for a given CaT value. A sample of 144 GCs in NGC 1407 suitable for stellar population analysis were obtained in the spectral region near the CaT using DEIMOS on Keck. The metallicity distribution for this sample was obtained based on the empirically determined conversion of AZ88 from CaT$_\mathrm{AZ88}$ index values. Several unexpected results were obtained, the most notable of which is the identical CaT index values for the brightest blue and red GCs. Even though the bright red and blue GCs are well separated in color space, the CaT measurements suggest that they have a similar metallicity. We show that this result is independent of the index measurement method used since the average raw spectra for the brightest blue and red GCs themselves are nearly identical. Integrated light spectra of nearby (Local Group) resolved GCs with independently measured ages, metallicities and HB morphologies determined via color-magnitude diagrams are essential to determine what effects metallicity, age, HB stars or blue straggler stars have on the integrated light spectra of GCs in the near-infrared. Such a dataset would also help our understanding of these phenomena and may allow for the calibration of the CaT as a metallicity indicator for extragalactic GCs. \section*{Acknowledgements} We thank the referee for his/her constructive comments. We would like to thank Javier Cenarro, Alan Brito, George Hau and Aaron Romanowsky for helpful insight and discussions. We are grateful to J. Trevor Mendel for help with the template fitting procedure. CF thanks the Anglo-Australian Observatory for financial support in the form of a graduate top-up scholarship. DAF and RNP thank the ARC for financial support. This material is based upon work supported by the National Science Foundation under Grant AST-0507729. Support for this work was provided by NASA through Hubble Fellowship grant \#HF-01214.01 awarded by Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., for NASA, under contract NAS 5-26555. The data presented herein were obtained at the W.M. Keck Observatory, which is operated as a scientific partnership among the California Institute of Technology, the University of California and the National Aeronautics and Space Administration. The Observatory was made possible by the generous financial support of the W.M. Keck Foundation. The analysis pipeline used to reduce the DEIMOS data was developed at UC Berkeley with support from NSF grant AST-0071048. Based in part on data collected at Subaru Telescope, which is operated by the National Astronomical Observatory of Japan. \begin{appendix} \section{Error determination and sample selection}\label{Appendix:errors} We use the method described in Cardiel et al. (1998) and in appendix A2 of C01 to compute the errors on the raw indices (CaT$_{\rm AZ88}$). For the CaT index measured on the template fitted spectra it is difficult to assess the errors in our measurements and hence on the inferred metallicity exactly. This is because the template fitting procedure has modified the variance on each pixel in a nontrivial manner. While the fitting has essentially extrapolated across skylines, thereby minimizing their effect on the final index measurements, it is unclear how the fits themselves may be affected by skyline residuals and noise. Adding to this is the possible introduction of errors arising from the continuum fitting step. For this reason, we model the error on our CaT measurements using Monte Carlo methods similar to that used by Emsellem et al. (2004) and Weijmans et al. (2009). We first measure the `exact' CaT index value using our index measurement method (template and continuum fitting applied) on V03 SSP model spectra. These are then compared to CaT values measured on the same fitted spectra with added Poisson noise. The errors as a function of signal-to-noise shown in Figure \ref{fig:SN_deltaCaT} are the average difference between the two measurements. We find no trend between the errors and metallicity. The method detailed above does not describe our CaT errors perfectly. Nevertheless, it follows our index computation scheme closely and should be a reasonable error estimate. This is supported by the fact that the scatter in the CaT and [Fe/H] distributions is similar to the size of the error bars. There is one caveat with these error estimates: the effect of skyline residuals and other non-Poisson noise sources are not included. We suspect the presence of skyline residuals have a small effect on the errors since skyline contaminated regions are excluded by the \textsc{pPXF} routine. Based on the size of the errors we removed spectra with raw counts less than 80 (S/N$\sim9$ \AA$^{-1}$) from our sample, which corresponds to roughly an $i$-band magnitude of 22.0 (see Figure \ref{fig:i_SN}). This yields a maximum CaT error of 0.5 \AA \space and a maximum [Fe/H] error of roughly 0.25 dex. This left a sample of 144 spectra, sufficient for a statistical analysis. \begin{figure} \epsscale{0.7} \plotone{fa1.eps} \caption{Errors on the CaT and [Fe/H] as a function of signal-to-noise per \AA\space (S/N). The stars show the average difference between the CaT measured directly on the SSP model spectrum and that measured on the same spectrum but with added Poisson noise. The solid line is a fit to these data points. The top abcissa shows the number counts expected if the spectra contain Poisson noise only. Our observations range from $9\le$S/N$\lesssim60$}\label{fig:SN_deltaCaT} \end{figure} \begin{figure} \epsscale{0.7} \plotone{fa2.eps} \caption{Variation of the average signal-to-noise per \AA\space (S/N) as a function of $i$-band magnitude for the spectroscopically confirmed GCs. The shaded grey area shows the sample of 144 GCs selected for the CaT analysis.}\label{fig:i_SN} \end{figure} \section{Other weak features}\label{sec:SmallLines} In this section, we explore other weak spectral features that are not clearly visible in the raw spectra but can be measured in the fitted spectra only. While the relative strength of these weak lines may be an artifact of the fitting technique, we find that there are some interesting characteristics that may help interpret our measured CaT distribution. The first interesting feature is the presence of the Pa12 line in a non-negligible number of our fitted spectra. An example of this is seen in Figure \ref{fig:indices}. It is the broad feature centered at 8751 \AA\space and it is overlaid by a multitude of smaller features. The Pa12 line is part of the Paschen series of which 3 features overlap with the CaT features, namely the Pa13, Pa15, and Pa16 lines. SSP model spectra do not show such Paschen lines with corresponding intermediate colors for ages $\gtrsim2.0$ Gyrs although one must keep in mind the increased uncertainty of the models at young ages. The hollow symbols in Figure \ref{fig:BICaT} show the GCs with measured Pa12 $\ge1.65$ \AA. Most of the data that are scattered to higher CaT index values are those with possibly large Pa12 absorption. Indeed, roughly 20 \% of the fitted spectra at intermediate colors/metallicity have Pa12 $\ge1.65$. The inferred average Pa12 index value for the bright blue GCs is 0.3 \AA\space greater than that of the bright red GCs indicating that the blue GCs may contain a higher degree of Paschen contamination. Unfortunately, it is not possible with the current data to positively confirm the presence of Paschen line contamination. We find that our fitted spectra also contain tentative evidence for the presence of several other `weak lines' (i.e. Ti\thinspace\textsc{i} $\lambda$8436\AA, Fe\thinspace\textsc{i} $\lambda$8514\AA\space and Fe\thinspace\textsc{i} $\lambda$8689\AA, see Figure \ref{fig:indices}). As with the Pa12 feature, these are not seen in our raw spectra but they appear in the fitted spectra. These lines are metallicity sensitive lines and could thus serve as potential metallicity indicators provided an appropriate calibration and signal-to-noise. As a matter of fact, other groups have made use of various spectral features in the region of the CaT to measure stellar metallicities using high signal-to-noise DEIMOS spectra (e.g.,Kirby et al. 2008; Shetrone et al. 2009). Figures \ref{fig:small_CaT} and \ref{fig:BI_small} show that the sum of the equivalent widths of the weak lines correlate with both CaT and color albeit with large scatter. The tighter correlation between the weak lines and CaT (Figure \ref{fig:small_CaT}) suggests that both trace metallicity similarly. On the other hand, while bimodality is not readily visible in the distribution of CaT, it is clear what the distribution of the sum of the weak lines is bimodal. Indeed, the KMM test concludes that a bimodal distribution is preferred over a unimodal at the 97 \% confidence with 65 and 35 \% of the GCs being blue and red, respectively. This is in contrast to their color proportion that are 40 and 60 \% for the blue and red GCs, respectively. When fitting our spectra we do not directly fit the small lines, which are not seen in the raw spectra. The templates themselves contain information about the relative ratios of these lines to CaT for a given metallicity such that the distribution for the sum of the weak lines displays the expected bimodality. We speculate that while the CaT features may be saturated at high metallicity, the weak lines could still be on the growth curve and their sum a better proxy for metallicity. We are very cautious however as these features are solely measured as a result of the template fitting method and thus the inferred bimodality could be an artifact. On the other hand, if it were an artifact it is not clear how the fitting method could conspire to create a clearer (i.e. with higher confidence) bimodality than seen in the CaT distribution. \begin{figure} \epsscale{0.7} \plotone{fb1.eps} \caption{Correlation between the CaT and the sum of the equivalent widths of the weak lines (Ti\thinspace\textsc{i} $\lambda$8436\AA, Fe\thinspace\textsc{i} $\lambda$8514\AA\space and Fe\thinspace\textsc{i} $\lambda$8689\AA), which are likely sensitive to metallicity. Filled and hollow squares correspond to Pa12 $< 1.65$ \AA\space and Pa12 $\ge 1.65$ \AA, respectively, with relative sizes proportional to the the signal-to-noise in the raw spectrum. The overall trend suggests that both CaT and the weak lines are tracing metallicity.}\label{fig:small_CaT} \end{figure} \begin{figure} \epsscale{0.7} \plotone{fb2.eps} \caption{Correlation between $(B-I)_0$ color and the sum of the equivalent widths of the three weak lines. Histograms of each variable show that both color and the equivalent widths of the weak lines are bimodal.}\label{fig:BI_small} \end{figure} \end{appendix}
\section{Complexity depends on vocabulary} \section{Introduction} While sparse polynomials are a natural data structure for human beings (who writes $x^{10}+0x^9+0x^8+0x^7+0x^6+0x^5+0x^4+0x^3+0x^2+0x^1-1$?) and computer algebra systems, algorithms to do more than add and multiply are scarce on the ground, and most texts slip silently from considering sparse polynomials to considering dense ones \cite{DavenportCarette2010}. This is partly because of the existence of examples showing that the output can be exponentially larger than the input, and hence ``nothing can be done''. We contend that these examples are basically all cases of the cyclotomic polynomials in disguise, and that, by admitting these to the {\it output\/} language, as Schinzel's $K$-operator \cite{Schinzel1982} effectively does, these examples cease to be absolute barriers to efficient algorithms. Cyclotomic factors can often be recognised relatively efficiently \cite{BradfordDavenport1989}, though the worst-case is NP-hard from this result. \def\cite[Theorem 6.1]{Plaisted1977}{\cite[Theorem 6.1]{Plaisted1977}} \begin{theorem}[\cite[Theorem 6.1]{Plaisted1977}]It is NP-hard to solve the problem, given a polynomial $p(x)\in\Z[x]$, to determine if $p$ has a root $r$ of modulus 1.\label{thm:Plaisted3} \end{theorem} This is a paradigmatic example of a more general thesis: solving problems in computer algebra requires the concurrent design of the most appropriate vocabulary and algorithms which are polynomial {\it in the size of the output\/} so encoded. Naturally, unconstrained multiplication of new vocabulary is not a viable solution, and a methodology for costing this was proposed in \cite{Carette2004}. A common problem in computer algebra is ``factorize this polynomial''. The algorithms commonly used first compute factorizations $p$-adically, and then deduce the ``true'' factorization over $\Z$. The traditional approaches \cite{Zassenhaus1969} are theoretically exponential in the number of $p$-adic factors, though in practice the exponential aspect can be ``controlled'' \cite{Abbottetal2000a}. Polynomial-time (in the degree, and {\it a fortiori\/} in the number of $p$-adic factors) algorithms are known \cite{Lenstraetal1982}, but in practice tend to be slower. The most recent progress is in \cite{vanHoeij2002}, whose algorithm is faster in practice, and the deduction phase is heuristically polynomial time in the number of $p$-adic factors. Of course, for a sparse polynomial such as $x^n-1$, the size of the polynomial is $O(\log n)$, and so an algorithm polynomial in $n$ is still exponential in the size of the input. Are there algorithms which are polynomial in the size of the input? If the output represents the factors as expanded polynomials, this is impossible. There is however a conjecture that the only cases which cause exponential blowups are cyclotomic factors -- we will return to this later. \begin{notation}We define the following for a polynomial $$f=\sum_{i=0}^na_ix^i=a_n\prod_{i=1}^n(x-\alpha_i):$$ \end{notation} \begin{description} \item[$\#f $]the number of non-zero terms in $f$, $|\{i: a_i\ne0\}|$; \item[${f_{-1}}$]$=x^nf(1/x)=\sum_{i=0}^na_{n-i}x^i$; \item[$f_-$]$=f(-x)=\sum_{i=0}^na_i(-x)^i$; \item[$f_e$]the even part of $f$, $\sum_{i=0}^{n/2}a_{2i}x^i$; \item[$f_o$]the odd part of $f$, $\sum_{i=0}^{n/2}a_{2i+1}x^i$; \item[$f_2$]the root-square, or Graeffe\footnote{There are various conventions in the literature as to how one handles the $\pm$ arising from the parity of $n$.}, of $f$, $\pm a_n^2\prod_{i=1}^n(x-\alpha_i^2)= f_e^2-xf_o^2$; \item[$||f||_\infty$]$=\max_{i=0}^n|a_i|$; \item[$||f||_2$]$=\sqrt{\sum_{i=0}^n a_i^2}$; \item[$||f||_1$]$={\sum_{i=0}^n |a_i|}$; \item[$M(f)$](the Mahler measure of $f$) $=|a_n|\prod_{i=1}^n\max(1,|\alpha_i|)$. \end{description} \begin{table*}[t] \caption{Large coefficients in $\Phi_k$}\label{tab:phi} \begin{tabular}{lrrrrrrrrrrrrr} $|a_i|$&2&3&4&5&6&7&8=9&14&23&25&27&59&359\\ first $\Phi_k$&105&385&1365&1785&2805&3135&6545&10465&11305&17225&20615&26565 &40755\\ $\phi(k)$&48&240&576&768&1280&1440&3840&6336&6912&10752&12960& 10560&17280\\ \end{tabular} \hfil\break [A larger version, independently computed, is in \cite{ArnoldMonagan2008}.] \end{table*} \section{Cyclotomic Polynomials} \begin{notation} Let $d(n)$ denote the number of divisors of the number $n$ (including 1 and $n$ itself). \end{notation} \begin{theorem}\label{thm:Wigert} It is known \cite{Wigert1907} that \begin{equation} d(n)\le n^{(\log 2+o(1))/\log\log n}. \end{equation} \end{theorem} However, we should note the caveats about the distribution of $d(n)$ given at \cite[Theorem 432]{HardyWright1979}, in particular that it has {\it average\/} order $\log n$ but {\it normal\/} order roughly $(\log n)^{\log 2}$. \begin{definition} We will say that a polynomial is {\em cyclotomic\/}, if all its roots are roots of unity. \rm Many authors reserve this for irreducible polynomials, but we will explicitly say ``irreducible'' when we need to. \end{definition} \begin{notation} Let $\Phi_k$ be the $k$-th irreducible cyclotomic polynomial: \begin{equation} \Phi_k(x)=\prod_{\gcd(j,k)=1}(x-e^{2\pi ij/k}). \end{equation} We denote by $C_n$ the cyclotomic polynomial with all $n$-th roots of unity, i.e. $C_n = x^n-1$. \end{notation} We should note that it is {\it not\/} the case that the coefficients of $\Phi_k$ are 0 or $\pm1$. The first counterexample is $\Phi_{105}$, which contains the terms $-2x^7$ and $-2x^{41}$. $\Phi_{385}$ contains the terms $-3x^{120}$ and $-3x^{121}$ The growth rate is in fact greater than one might expect, and $\Phi_{15015}$ has terms of $23\,{x}^{2294}$ and $23\,{x}^{3466}$. $15015=3\cdot5\cdot7\cdot11\cdot13$, and this looks like the recipe (confirmed in \cite{Batemanetal1984}) to make large values\footnote{It {\it does\/} give us (exactly) 500 at $255255=3\cdot5\cdot7\cdot11\cdot13\cdot17$.}, but in fact 23 is first attained at $11305=5\cdot7\cdot17\cdot19$, as shown in table \ref{tab:phi}. We note the spectacular leap at $40755=3\cdot5\cdot11\cdot13\cdot19$, which is the largest coefficient up to $k=80,000$. \cite[Table 3]{ArnoldMonagan2008} shows more such leaps for (much) larger $n$. \begin{theorem}\label{thm:vaughan} \cite[Theorem 1]{Vaughan1974} shows that, for infinitely many $n$, \begin{equation}\label{eq:Phi} \log||\Phi_n||_\infty>\exp\left(\frac{(\log 2)(\log n)}{\log\log n}\right), \end{equation} \end{theorem} and indeed this is precisely the right order of (worst-case) growth \cite{Bateman1949}, perhaps better expressed as \begin{equation}\label{eq:Phigrowth} \limsup_{n\rightarrow\infty}\frac{||\Phi_n||_\infty}{\log n/\log \log n}=\log 2. \end{equation} \begin{proposition}$x^n-1=\prod_{d|n}\Phi_d(x)$, and these factors are irreducible. \end{proposition} \begin{proposition}\label{prop:Mobius} $\Phi_n(x)=\prod_{d|n}(x^d-1)^{\mu(n/d)}$, where $\mu$ is the M\"obius function. \end{proposition} \begin{proposition}$x^n-1$ has $d(n)$ irreducible factors. \end{proposition} Cyclotomic polynomials are the bugbear of anyone who tries to deal with sparse polynomials. \begin{example}\label{ex:factor} Asking for the factorization, or even the degrees of the factors, of $x^n-1$ is tantamount to factoring $n$, since for every prime $p$ dividing $n$, there is an $x^{p-1}+\cdots$ in the factorization of $x^n-1$. \end{example} \begin{example}\label{ex:phifact} Similarly, asking for the degree of $\Phi_k$ is, if $k$ is $p\cdot q$ ($p$, $q$ distinct primes), tantamount to factoring $k$, since $\phi(k)=(p-1)(q-1)$ and so $$ p,q = \frac12\left(k+1-\phi(k)\pm\sqrt{k^2-2k-2k\phi(k)+(\phi(k)-1)^2}\right). $$ \end{example} Cyclotomic polynomials are frequently used as examples. \begin{example}\cite[p. 185]{vanHoeij2002} gives this example $$ {x}^{128}-{x}^{112}+{x}^{80}-{x}^{64}+{x}^{48}-{x}^{16}+1, $$ and states that his algorithm sped up Maple by a factor of 500 on this example. From a cyclotomic-aware point of view, such as \cite{BradfordDavenport1989}, this polynomial is easy. Four applications of Graeffe's root-squaring process show (as is obvious to the eye) that this is $f(x^{16})$ where $$ f(x) = {x}^{8}-{x}^{7}+{x}^{5}-{x}^{4}+{x}^{3}-x+1. $$ Another application takes $f$ to itself, and hence $f$, and so the original polynomial, is cyclotomic. If $\alpha$ is a root of $f$, $\alpha^{15}=1$, so $f=\Phi_{15}$, and the original polynomial is $$ \Phi_{15}\Phi_{30}\Phi_{60}\Phi_{120}\Phi_{240}. $$ \end{example} \begin{example} If $p$ is prime, $f=x^p-1$, then $\#f=2$ but $f=(x-1)(x^{p-1}+\cdots+1)$: two factors with 2 and $p$ terms respectively. \end{example} \begin{example}\label{ex:twofact} If $p$, $q$ are distinct primes, $f=(x^p-1)(x^q-1)$, then $\#f=4$ but $f=(x-1)^2(x^{p-1}+\cdots+1)(x^{q-1}+\cdots+1)$. The square-free decomposition of $f$ is therefore a repeated factor with 2 terms and a factor $g$ with $pq-p-q+2$ terms respectively. The largest coefficient in $g$ is $\min\{p,q\}$, and $||g||_1=pq$. \end{example} Obviously, a square-free decomposition of $f$ was a bad idea in this case: however previously-proposed algorithms, e.g. \cite[p. 69]{BeukersSmyth2002} tend to do this. \par It could be argued that the problem in this case is the `cofactor', but life is not that simple. \begin{example} If $p$, $q$ are distinct primes, $f=(x^p-1)^2(x^q-1)$, then $\#f=6$ but the square-free factorization is $$ (x-1)^3\left(x^{p-1}+\cdots+1\right)^2\left(x^{q-1}+\cdots+1\right), $$ and we are forced to write out the large squared factor. The largest coefficient of $\left(x^{p-1}+\cdots+1\right)^2$ is $p$, so we had also better not compute it and then take its square root. \end{example} It is the contention of this paper that all these difficulties {\it except the first\/} are caused by an inadequate vocabulary: the first seems to be intrinsic, in the fact the factorization of numbers can be encoded as a problem of factorization of polynomials. All we can do is recognise the fact. \section{Representational Complexity}\label{sec:RC} Information theory, whether through the guise of Kolmogorov Complexity \cite{LiVitanyi} or Minimum Description Length \cite{MDLbook}, tell us that \emph{good} representations of structured objects are two-part codes: a model and an encoding of data using that model. In other words, the proper ``length'' of an object consists in counting the length of the representation of the model as well as the representation of the data encoded using this model. \par In \cite{Carette2004}, these results from information theory are rephrased so as to apply more directly to Computer Algebra Systems, and applications to simplification are outlined. The basic result is that for large enough structured expressions, it is \emph{always} worthwhile to first formalize the ``structure'', and then encode the data in such a way as to abstract out that structure. Note that, if model extensions are not allowed, then simplification reduces simply to length reduction. It is exactly the confusion between issues of the (background) model-class and its use in model reduction which caused Moses \cite{Moses1971a} to argue that ``simplification'' was impossible to formalize. \par When tackling a particular situation, \cite{Carette2004} boils down to finding the right \emph{vocabulary} in which to express ones' result. In some cases, the right vocabulary is somewhat counter-intuitive. For example, in the case of algebraic numbers, it was long ago discovered that using minimal polynomials to ``encode'' an algebraic number was best -- although this can seem puzzling in the setting of ``solving'' a polynomial, as then the answer to the problem is just an encoding of the question. We will return to this issue later. \par For the particular case of factoring of polynomials, what does this tell us? All the theoretical results point in the same direction: cyclotomic polynomials are somehow ``special cases'', especially when one is factoring sparse polynomials. Conventional wisdom already tells us that both dense and sparse polynomials are useful model classes, and that we should have both at our disposal. The mathematical theory of factoring polynomials (as outlined in the rest of this paper) informs us that cyclotomics are undeniably part of the domain of discourse. Combining these together tells us that they should also be part of our model classes. The only remaining question then is whether adding this particular vocabulary actually leads to a \emph{simplification}. To evaluate this, we need to actually display some data structures designed with this new vocabulary, and then evaluate if we have made any real gains. \section{Data Structures}\label{sec:DS} Since we will be arguing on the size of data structures, we will need to define our data representations. The precise details might vary, though in practice the conclusions will not. For definiteness, we describe our choices according to the {\it unaligned\/} packed encoding of ASN.1 \cite{ITU2002}: note that their {\tt SEQUENCE} is what C programmers would think of as {\tt struct}. Our encodings are intended to be practical, though we ignore issues of alignment to word boundaries, and indeed a number of $\lceil\ldots\rceil$ operators are also omitted. \par In theory, one needs to have arbitrary sized data fields, which means one needs fields for the length of the size fields. To avoid this, we will assume that $N$ and $K$ are global parameters for the size of the object, bounding the degrees and the size of the coefficients. Since a polynomial's factors always have smaller degree, we can assume $N=\log_2n$. It is not so easy for the coefficients \cite{Mignotte1981a}, but we will assume a {\it single\/} field of $K$ bits associated with each outermost data structure, giving the size $k$ of all the coefficients stored in that structure. Since there is one of these, we can ignore its cost. \par We next give explicit representations for dense polynomials, sparse polynomials, factored polynomials, $\Phi$-aware factorizations and $C$-aware factorizations (explained fully below). \subsection{A single dense polynomial} We choose a dense representation with a uniform size bound for all the coefficients. A single dense polynomial of degree $n$ requires $\log_2 n$ bits to represent the degree, and then there are $n+1$ coefficients. Hence if each of them requires $k$ bits, we need $\log_2 k$ bits for the telling us this, and then the coefficients require $k+1$ bits (including sign). \begin{equation} (k+1)(n+1)+\log_2 k+\log_2 n \label{Dense1} \end{equation} In pseudocode\footnote{Essentially ASN.1, except that we allow ourselves to write mathematics, enclosed in boxes, in the pseudocode.}, this might be represented as follows. \begin{lstlisting}[mathescape] DensePoly ::= SEQUENCE { Degree INTEGER ($\psom{0\ldots 2^N-1}$), k INTEGER ($\psom{0\ldots 2^K-1}$), Coefficients SEQUENCE { INTEGER ($\psom{-2^k\ldots 2^k-1}$) } SIZE ($\psom{\hbox{\tt Degree}+1}$) } \end{lstlisting} \subsection{A single sparse polynomial} We choose a sparse representation with a uniform size bound for all the coefficients. Furthermore, we assume that there are $t$ non-zero coefficients, i.e. $t$ terms to be represented, and $t$ is bounded by the same bound as the degree. A single sparse term from a polynomial of degree $n$ requires $\log_2 n$ bits to represent the degree. Hence the total space is given by \begin{equation} \log_2 n + t(k+ 1+\log_2 n). \label{Sparse1} \end{equation} In pseudocode, this might be represented as \begin{lstlisting}[mathescape] SparsePoly ::= SEQUENCE { TermCount INTEGER ($\psom{0\ldots 2^N-1}$), Terms SEQUENCE { Degree INTEGER ($\psom{0\ldots 2^N-1}$), Coefficient INTEGER ($\psom{-2^k\ldots 2^k-1}$) } SIZE (TermCount) } \end{lstlisting} Using a Horner scheme might save a few more bits in the representation of the exponent, but rarely appreciably so. \subsection{Representing Factorizations}\label{sec:mult} We will use the same structure for square-free or complete factorizations: the number of (distinct) factors followed by pairs (multiplicity, factor). In pseudocode, this might be represented as \begin{lstlisting}[mathescape] Factorization ::= SEQUENCE { FactorCount INTEGER ($\psom{0\ldots 2^N-1}$), Factors SEQUENCE { Multiplicity INTEGER ($\psom{0\ldots 2^N-1}$), Factor $\psom{Poly}$ } SIZE (FactorCount) } \end{lstlisting} where $Poly$ is one of \verb+DensePoly+ or \verb+SparsePoly+. Hence with $f$ factors, the overhead (i.e. the cost over and above that of storing the distinct factors themselves) is \begin{equation} (f+1)\log_2 n. \end{equation} \subsection{Representing $\Phi$-aware Factorizations}\label{sec:pmult} We will use the same structure for square-free or complete factorizations: the number of (distinct) factors followed by pairs (multiplicity, factor). However, we do this twice: once for the factors that are $\Phi_k$, and once for those that are not. The factors that {\it are\/} $\Phi_k$ are stored as $k$ followed by\footnote{The inclusion of $\phi(k)$ avoids the problem in example \ref{ex:phifact}.} $\phi(k)$. In pseudocode, this might be represented as \begin{lstlisting}[mathescape] Factorization ::= SEQUENCE { PhiFactorCount INTEGER ($\psom{0\ldots 2^N-1}$), PhiFactors SEQUENCE { Multiplicity INTEGER ($\psom{0\ldots 2^N-1}$), k INTEGER ($\psom{0\ldots 2^N-1}$), Degree INTEGER ($\psom{0\ldots 2^N-1}$), } SIZE (PhiFactorCount), FactorCount INTEGER ($\psom{0\ldots 2^N-1}$), Factors SEQUENCE { Multiplicity INTEGER ($\psom{0\ldots 2^N-1}$), Factor $\psom{Poly}$ } SIZE (FactorCount) } \end{lstlisting} where $Poly$ is one of \verb+DensePoly+ or \verb+SparsePoly+. Hence with $l$ $\Phi_k$ factors, the cost of storing them is $(3l+1)\log n$. {\it Any\/} $\Phi_k$ stored this way is cheaper than in any of the previous representation (dense, sparse or factored), so the worst case is when there are no $\Phi_k$ in the factorization, when the overhead is merely the one field \verb+PhiFactorCount+. \subsection{Representing $C$-aware Factorizations}\label{sec:cmult} Instead of storing the $\Phi_k$, we could store the complete cyclotomic polynomials $x^k-1$. Again, we will use the same structure for square-free or complete factorizations: the number of (distinct) factors followed by pairs (multiplicity, factor). Also in this case, we do this twice: once for the factors that are of the form $x^k-1$, and once for those that are not. One might think to store the factors that {\it are\/} $x^k-1$ simply as $k$. However, as pointed out in example \ref{ex:factor}, this will not allow us to answer questions such as ``how many factors'' in a reasonable time. Hence we store $k$ followed by its prime factorization. \par We should also note that Proposition \ref{prop:Mobius} means that we now need {\it negative\/} multiplicities as well. In pseudocode, this might be represented as \begin{lstlisting}[mathescape] Factorization ::= SEQUENCE { CFactorCount INTEGER ($\psom{0\ldots 2^N-1}$), CFactors SEQUENCE { Multiplicity INTEGER ($\psom{-2^N\ldots 2^N-1}$), Degree INTEGER ($\psom{0\ldots 2^N-1}$), NumFactors INTEGER ($\psom{0\ldots 2^N-1}$), KFactor SEQUENCE { INTEGER ($\psom{0\ldots 2^N-1}$) } SIZE (NumFactors), } SIZE (CFactorCount), FactorCount INTEGER ($\psom{0\ldots 2^N-1}$), Factors SEQUENCE { Multiplicity INTEGER ($\psom{0\ldots 2^N-1}$), Factor $\psom{Poly}$ } SIZE (FactorCount) } \end{lstlisting} where $Poly$ is one of \verb+DensePoly+ or \verb+SparsePoly+. Hence with $f$ factors of $x^k-1$ involved, the overhead (i.e. the cost over and above that of storing the distinct factors themselves) is \begin{equation} (f+1)\log_2 n. \end{equation} \par An alternative formulation might store the factors of $k$ with multiplicity: there does not seem to be a great deal to choose between them. In either case, we should note that asking for the {\it number\/} of irreducible factors is no longer trivial, since the number of irreducible factors corresponding to a single $x^k-1$ is $d(k)$. Furthermore, since we are allowed negative exponents, representing $$x^k+1 :=\left(x^{2k}-1\right)/\left(x^k-1\right),$$ we have to note that not all the irreducible factors of $x^{2k}-1$ are actually factors of the left-hand side. Nevertheless, since we have stored the prime factorization of $k$, the problem is efficiently soluble (certainly polynomial time in the size of the representation). \begin{table*}[th] \caption{\label{table:cyc}Representing $x^n-1=\prod_{d|n}\Phi_d(x)$} \begin{tabular}{rrrr} Representation&Fully expanded&Square-free factorization&Factored\\ Dense&$2(n+1)+\log_2 n$&same as factored& $n^{1+\frac{\log 2}{\log\log n}}\log_2e$ \\ Sparse&$3\log_2 n$&same as factored& $n^{1+\frac{\log 2}{\log\log n}}\log_2e$ \\ $\Phi_k$&$3\log_2 n$&same as factored&$(2d(n)+1)\log_2 n$ \\ $x^k-1$&$3\log_2 n$&same as factored&$\log_2 n$ \\ \end{tabular} \end{table*} \section{Representing some cyclotomic polynomials}\label{sec:cycrep} \begin{table*}[t] \caption{Representing $(x^p-1)(x^q-1)=(x-1)^2\Phi_p(x)\Phi_q(x)$ with $n=p+q$\label{table:cyc2}} \begin{tabular}{rrrr} Representation&Fully expanded&Square-free factorization&Factored\\ Dense&$2(n+1)+\log_2 n$&$(1+\log_2 n)(n+2)+4\log_2 n$&$2(n+3)+6\log_2 n$ \\ Sparse&$4\log_2 n$&$(2n+2)\log_2 n$&$(n+10)\log_2n$ \\ $\Phi_k$&$4\log_2 n$&$6\log_2 n$&$6\log_2 n$ \\ $x^k-1$&$4\log_2 n$&$2\log_2 n$&$2\log_2 n$ \\ \end{tabular} \end{table*} We now use each of our representations and compute the size of the results. \subsection{Factorization of $x^n-1$}\label{sec:dense1} The sizes of the expanded $x^n-1$ polynomial in dense, and sparse encodings are obvious. The $\Phi$-aware and $C$-aware versions are within an additive constant of \verb+DensePoly+ / \verb+SparsePoly+. For definiteness, we will use \verb+SparsePoly+ based counts. \par To understand the size of the factored forms, we need to study their sizes a little more closely. In this case, the factored form and the square-free form coincide. There are $d(n)$ factors, of total degree $n$, hence $n+d(n)$ terms. By Theorem \ref{thm:vaughan}, we can bound\footnote{In theory, not all the factors {\it can\/} have coefficients this large, but the gain from exploiting this is relatively small.} the size of the coefficients as $\log_2e$ times the right-hand side of (\ref{eq:Phi}), {\it viz.\/} $\frac{\log n}{\log\log n}$, to get the size of a dense polynomial to be $$ (n+d(n))\log_2e\exp\left(\frac{(\log 2)(\log n)}{\log\log n}\right)+ d(n)\log_2n $$ \begin{equation}\label{eq:densephi} \approx n^{1+\frac{\log 2}{\log\log n}}\log_2e. \end{equation} \par {\it In general\/}\footnote{$x^{2^k}-1$ is an obvious counter-example.}, the factors will be essentially dense, so a sparse encoding will save nothing, but have to pay for the cost of storing the degrees with each coefficient, adding $(n+d(n))\log_2n$, to give $$ (n+d(n))\left(\log_2n+\log_2e\exp\left(\frac{(\log 2)(\log n)}{\log\log n}\right)\right)+ d(n)\log_2n $$ \begin{equation}\label{eq:sparsephi} \approx n^{1+\frac{\log 2}{\log\log n}}\log_2e. \end{equation} We should note that the asymptotically dominant term is the coefficient storage in this model, which is contrary to intuition, and even the experimental data in table \ref{tab:phi}, but this merely shows that the asymptotics will take time to be visible. \par The results of this section are summarised in Table \ref{table:cyc}. \subsection{A square-free factorization}\label{sec:sqfr1} Let use now consider $(x^p-1)(x^q-1)$, with $p,q$ distinct primes. The ``fully expanded'' versions are again obvious. The square-free factorization of $(x^p-1)(x^q-1)=(x-1)^2\Phi_p(x)\Phi_q(x)$ involves multiplying out $\Phi_p(x)\Phi_q(x)$. This gives us coefficients of size $O(n)$, in fact $n/2$ assuming that $p$, $q$ are balanced, taking $(\log_2 n)-1$ bits to represent the magnitude\footnote{In this case, they are all positive, but we can't count on this in general.}. \par In the factored representation, we have three factors, of degrees 1, $p-1$ and $q-1$, i.e. total degree $n-1$. All coefficients are bounded by 1. Hence the total is \begin{equation}\label{eq:sparsesqfr} (n-1+3)(\log_2n+2) + (3\cdot2+2)\log_2n \approx (n+10)\log_2n. \end{equation} \par The results of this section are summarised in Table \ref{table:cyc2}. \section{Implementation notes} \subsection{Cyclotomic-free} It is important to review the encodings of the previous section and notice that for cyclotomic-free cases, these encodings involve constant overhead, independent of the degree and of the number of factors. In fact, by using a bit or two in a header word (which modern computer algebra systems always seem to use in their internal representations), one can choose between these encodings as necessary. In other words, cyclotomic-free polynomials do not have to bear any extra representation cost for this vocabulary extension. \par We can also construct various mixed cases, in other words sparse polynomials which factor into a cyclotomic part and a small dense cofactor. The difference in encoding cost is correspondingly mixed, although the end result is similar: adding cyclotomics asymptotically wins. \subsection{Which to choose?} We have posited two encodings for ``cyclotomic-aware'' representations of factorizations: one in terms of the irreducible cyclotomics $\Phi_k$ (section \ref{sec:pmult}) and one in terms of the `complete' cyclotomics $C_k=x^k-1$ (section \ref{sec:cmult}). Tables~\ref{table:cyc} and~\ref{table:cyc2} make it clear that adding cyclotomics to ones' vocabulary is certainly reprensentionally efficient. But which one should be used? We first summarize some the advantages and disadvantages. \begin{description} \item[Pro $\Phi_k$: clearly polynomial]In the $\Phi_k$-representation, the factorization of $x^{p+q}+x^p+x^q+1$ is $\Phi_2^2\Phi_{2p}\Phi_{2q}$, where as in the $C_k$-representation it is $C_{2p}C_p^{-1}C_{2q}C_q^{-1}$. Functions meant to extract information from products of polynomials (degrees, multiplicity, etc) still function `easily', whereas spotting that the $C_k$-representation is not square-free is cheap, but not `obvious'. \item[Anti $\Phi_k$: worst-case blowup]The last two entries of table \ref{table:cyc} show that, for highly composite $n$, representing $C_n$ in terms of $\Phi_k$ can require almost $n$ times as much, i.e. exponentially more, since the input has size $O(\log n)$, space, by Theorem \ref{thm:Wigert}. \item[Notes]The pragmatist would probably choose $\Phi$. The theoretician would be swayed by the complexity argument and want $C$. Possibly the best answer is to admit $\Phi$, and $C$ with an additional extension to the vocabulary as in (\ref{eq:Cphi}). \end{description} To further illustrate one particular difficulty, let us consider factoring of $C_{105}$. It factors into $$ {\Phi_{1}(x)}{\Phi_{3}(x)} {\Phi_{5}(x)} {\Phi_{7}(x)} {\Phi_{15}(x)} {\Phi_{21}(x)} {\Phi_{35}(x)} {\Phi_{105}(x)}. $$ But if all we have at our disposal is $C$, then the best we can do (which is still better than using sparse polynomials) is $$ C_1(x)\frac{C_3(x)}{C_1(x)}\frac{C_5(x)}{C_1(x)}\frac{C_7(x)}{C_1(x)} \frac{C_{15}(x)C_1(x)}{C_3(x)C_5(x)} \frac{C_{21}(x)C_1(x)}{C_3(x)C_7(x)}\times $$ $$ \frac{C_{35}(x)C_1(x)}{C_5(x)C_7(x)} \frac{C_{105}(x)C_{7}(x)C_{5}(x)C_3(x)}{C_{35}(x)C_{21}(x)C_{15}(x)C_1(x)}. $$ An even more succinct representation is \begin{equation}\label{eq:Cphi} \prod_{\displaystyle k=1\atop k|105}^{105} \Phi_k(x) \end{equation} at the cost of another vocabulary extension. This is why in the representation of $C$ in section~\ref{sec:cmult}, we list the factors of $n$, which allows us to recover this factorization relatively easily. This still means that in the $\Phi$-aware encoding, obtaining the number of factors, the degrees of each of the factors, or the multiplicity of each of the factors is straightforward, while for the $C$-aware encoding, these simple questions now require some (small) amount of computation. , and so forth. All of these questions remain just as easy to answer in the $\Phi$-aware encodings as they were before. However, for the $C$-aware encodings, these ``simple'' questions now require some actual computations to resolve. In other words, adding $\Phi$ to our vocabulary is a very minor change with clear efficiency gains, while adding $C$ is slightly disruptive but with even greater asymptotic efficiency gains. \section{Conclusion} In section~\ref{sec:cycrep} we have shown how some ``troublesome'' factorizations, i.e. examples \ref{ex:factor} and \ref{ex:twofact}, cease to consume inordinate space when the cyclotomics are represented explicitly. But what of arbitrary polynomials and their factorizations? \par The answer is that we do not know, but there are some tantalizing results. \cite{Filasetaetal2008} state that, provided $f$ is non-reciprocal ($f\ne\pm f_{-1}$), $f$ has a factor with at most $c_2(\#f,||f||_\infty)$ terms, independent of $n$, and quotes \cite{Schinzel1969} to say that if the polynomial has no reciprocal factors, then all irreducible factors have at most $c_2(\#f,||f||_\infty)$ terms. This is a significant step towards controlling the dependence of the {\it output size\/} on $n$, though it falls short of saying that is polynomial in the following ways: \begin{itemize} \item there is no guarantee that $c_2$ depends {\it polynomially\/} on $\#f$ or $\log||f||_\infty$, and indeed $c_2$ is still rather mysterious to the authors \cite[Challenge 4]{DavenportCarette2010}; \item nothing is said about the size of the coefficients, and all known technology makes them depend exponentially on $n$ (i.e. the output size depends linearly on $n$). \end{itemize} We note, though, that the {\it known\/} examples of such growth \cite[$\S3$]{Mignotte1981a}, depend on cyclotomic polynomials, so one could hope that the second problem does not occur in practice. \par We are convinced, and we hope to have convinced the reader, that, regardless of whether using cyclotomics ultimately makes the problem of sparse polynomial factorization more tractable, even if not guaranteed polynomial (see Theorem \ref{thm:Plaisted3}), in the input size, that they are most definitely worth having in the basic vocabulary used for the {\it output\/} of factoring. We strongly recommend that the specification for what it means to factor a polynomial be thus amended.
\section{Introduction} \label{sec:intro} We work throughout over a Henselian discrete valuation field $K$ with ring of integers ${\mathcal{O}_K}$. We fix a uniformiser $t$, a normalisation $\nu$, and write $k={\mathcal{O}_K}/t{\mathcal{O}_K}$. Set $S=\operatorname{Spec}{\mathcal{O}_K}$. Let $C$ be a smooth genus one curve over $K$. $D$ will denote a $K$-rational divisor on $C$ of degree $n$. If $n=1$, then $C(K)\ne\emptyset$. If $n\ge 2$, then the divisor class $[D]$ defines a morphism $C\to{\mathbb P}_K^{n-1}$. Let $R$ be a Dedekind domain. A {\em genus one equation of degree $n$} describes the pair $(C,D)$ and is defined as follows: If $n=1,$ a Weierstrass equation \begin{equation} y^2 + a_1 x y + a_3 y = x^3 + a_2 x^2 + a_4 x + a_6,\;a_i\in K.\end{equation} Two genus one equations of degree $1$ with coefficients in $R$ are {\em $R$-equivalent} if they are related by the substitutions: $x' = u^2x + r$ and $y' = u^3y + su^2x + t,$ where $r,s,t\in R,\;u\in R^*.$ We set $\det([u;r,s,t])=u^{-1}.$ If $n=2,$ a (generalised) binary quartic \begin{equation}y^2 + (\alpha_0 x^2 + \alpha_1 xz + \alpha_2z^2) y = a x^4 + b x^3z + c x^2z^2 + d xz^3 + ez^4.\end{equation}Two genus one equations of degree $2$ with coefficients in $R$ are {\em $R$-equivalent} if they are related by the substitutions: $x'=m_{11}x+m_{21}z,\;z'=m_{12}x+m_{22}z$ and $y'=\mu^{-1}y+r_0x^2+r_1xz+r_2z^2,$ where $\mu\in R^*,\; r_i\in R$ and $M=(m_{ij})\in \operatorname{GL}_2(R)$. We set $\det([\mu,(r_i),M])=\mu\det M.$ If $n=3,$ a ternary cubic \begin{equation}F(x,y,z)=a x^3 + b y^3 + c z^3 + a_2x^2 y + a_3x^2 z + b_1y^2x+ b_3 y^2z + c_1z^2x + c_2 z^2y + m x y z=0.\end{equation} Two genus one equations of degree $3$ with coefficients in $R$ are {\em $R$-equivalent} if they are related by multiplying by $\mu\in R^*$, then substituting $x'=m_{11}x+m_{21}y+m_{31}z,\;y'=m_{12}x+m_{22}y+m_{32}z$ and $z'=m_{13}x+m_{23}y+m_{33}z$, where $M=(m_{ij})\in \operatorname{GL}_3(R).$ Set $\det([\mu,M])=\mu\det M.$ If $n\ge4$, then $C$ is described by $l=n(n-3)/2$ quadratic forms in $n$ variables. Two genus one equations of degree $n$ with coefficients in $R$ are {\em $R$-equivalent} if they are related by the substitutions: $F'_i=m_{i1}F_1+m_{i2}F_2+\ldots + m_{il} F_l,\; i=1,\ldots,l,$ for $M=(m_{ij})\in\operatorname{GL}_l(R),$ and then $x'_j=\sum_{i=1}^nn_{ij}x_i$ for $N=(n_{ij})\in\operatorname{GL}_n(R).$ When $n=4$, set $\det([M,N])=\det M\det N.$ For $n\le4$, we associate invariants $c_{4,\phi}$, $c_{6,\phi}$ and {\em discriminant} $\Delta_{\phi}$ to a genus one equation $\phi$ of degree $n$ such that $\Delta_{\phi}=(c_{4,\phi}^3-c_{6,\phi}^2)/1728.$ Moreover, $\phi$ defines a smooth curve of genus one if and only if $\Delta_{\phi}\ne 0.$ The invariants $c_{4,\phi},c_{6,\phi}$ and $\Delta_{\phi}$ are of weights $r=4,6$ and $12$ respectively. In other words, if $F\in\{c_{4,\phi},c_{6,\phi},\Delta_{\phi}\}$, then $F\circ g= (\det g)^rF$ for every $g$ defined by an $R$-equivalence. These invariants have been known since the nineteenth century, and can be found in \cite{AKMMMP}. We scale these invariants according to \cite{FiInvariants}. \begin{Definition}\label{def:integral-minimal} A genus one equation $\phi$ of degree $n$, $n\le4$, with $\Delta_{\phi}\ne 0$ is \begin{myitemize} \item[(a)] {\em integral} if the defining polynomials have coefficients in ${\mathcal{O}_K}.$ \item[(b)] {\em minimal} if it is integral and $\nu(\Delta_{\phi})$ is minimal among all the valuations of the discriminants of the integral genus one equations $K$-equivalent to $\phi.$ \end{myitemize} \end{Definition} Producing integral genus one equations of degree $n$ with small coefficients has been a target for investigations. In order to obtain such genus one equations, we need to {\em reduce} and {\em minimize} them. By reducing genus one equations, we mean reducing the size of the coefficients by a unimodular linear change of coordinates, which does not change the invariants. To minimise genus one equations, we need to make the associated invariants smaller. Reduced and minimal genus one equations of degree $2$ appear as an essential part of the 2-descent algorithm described by Birch and Swinnerton-Dyer in \cite{swinnerton}. More recent treatment for the minimisation of genus one equations of degree 2, 3 and 4 can be found in \cite{CremonaStollMini}, \cite{Fiminimise} and \cite{WomackThesis} respectively. An algorithmic approach for minimising genus one equations of degree $n$, $n\le4,$ can be found in \cite{FiStCr}. This paper will be dedicated to the minimisation question. The main obstacle which holds back the existence of a neat solution for the minimsation question when $n$ is large is the difficulty of describing the rings of invariants associated to these genus one equations. In order to overcome this difficulty, we give an alternative definition for the minimality of genus one equations of degree $n$. If $\phi$ is an integral genus one equation of degree $n$, then it defines an $S$-scheme ${\mathcal C}$. We give criteria for ${\mathcal C}$ to be normal, and hence an $S$-model for its generic fiber, see Definition \ref{def:models} below. Our definition of minimality compares the minimal desingularisation of ${\mathcal C}$ to the minimal proper regular model of its generic fiber. This definition does not involve invariant theory of genus one curves. We prove that our definition agrees with Definition \ref{def:integral-minimal} when $n\le 4$. In a sequel to this paper, we will use our definition of minimality to count the number of minimal genus one equations of degree $n$ up to ${\mathcal{O}_K}$-equivalence in a given $K$-equivalence class, for $n\le 4$. Finally, we give a new proof for the following theorem (\cite{FiStCr}, Theorem 4.17). \begin{Theorem} \label{thm1} Let $F$ be a number field of class number $1$ with ring of integers $\mathcal{O}_F$. Let $C$ be a smooth genus one curve defined over $F$ by a genus one equation $\phi$ of degree $n$ for $n\le4$. Assume that $C(F_{\nu})\ne\emptyset$ for every completion $F_{\nu}$ of $F$. Let $E$ be the Jacobian elliptic curve of $C$ with minimal discriminant $\Delta$. Then $\phi$ is $F$-equivalent to an $\mathcal{O}_F$-integral genus one equation whose discriminant is $\Delta.$ \end{Theorem} \section{Models of genus one curves} \label{sec:models of genus one curves} \begin{Definition}\label{def:models} An {\em $S$-curve} is an integral, projective, flat, normal $S$-scheme $f:X\to S$ of dimension 2. The generic fiber of ${\mathcal C}$ will be denoted by ${\mathcal C}_K$ and its special fiber by ${\mathcal C}_k$. We define an {\em $S$-model} for a smooth curve $C$ over $K$ to be an $S$-curve ${\mathcal C}$ such that ${\mathcal C}_K$ is isomorphic to $C$. \end{Definition} \begin{Definition}\label{def:contraction} Let ${\mathcal C}$ be an $S$-curve. Let $(\Gamma_i)_{i\in I}$ be the family of irreducible components of ${\mathcal C}_k.$ For a strict subset $J\subset I$, a \emph{contraction} of the components $\Gamma_j,\;j\in J,$ in ${\mathcal C}$ consists of an $S$-morphism $u:{\mathcal C}\rightarrow {\mathcal C}^J $ of $S$-schemes such that\\ (a) For each $j\in J$, the image $u(\Gamma_j)$ consists of a single point $x_j\in {\mathcal C}^J$, and\\ (b) $u$ defines an isomorphism ${\mathcal C}-\bigcup_{j\in J}\Gamma_j\xrightarrow{\sim}{\mathcal C}^J-\bigcup_{j\in J}{x_j}.$ \end{Definition} Since ${\mathcal{O}_K}$ is Henselian, the contraction $u:{\mathcal C}\to{\mathcal C}^J$ of the components $\Gamma_j,\;j\in J,$ exists for any strict subset $J\subset I$. Moreover, the morphism $u$ is unique up to unique isomorphism, see (\cite{Liubook}, Theorem 8.3.36 and Proposition 8.3.28). The following theorem, Theorem $1$ of (\cite{Neron}, \S 6.7), describes the contraction morphism explicitly. \begin{Theorem} \label{th:contraction} Let ${\mathcal C}$ be an $S$-curve. Let $(\Gamma_i)_{i\in I}$ be the family of irreducible components of ${\mathcal C}_k.$ Let $D$ be a non-trivial effective Cartier divisor on ${\mathcal C}.$ Let $J$ be the set of all indices $j\in I$ such that $\operatorname{Supp} (D)\cap \Gamma_j=\emptyset.$ Then the canonical morphism $$u:{\mathcal C}\to{\mathcal C}^J:=\operatorname{Proj}(\bigoplus_{m=0}^{\infty}H^0({\mathcal C},\mathcal{O}_{{\mathcal C}}(m D)))$$ is the contraction of the components $\Gamma_j,j\in J,$ and ${\mathcal C}^J$ is an $S$-curve. \end{Theorem} Let $C$ be a smooth genus one curve over $K$. Assume that $C(K)\ne\emptyset$. Let $E$ be the Jacobian elliptic curve of $C$ with minimal proper regular model $E^{min}$. Since $C \cong _K E$, the minimal proper regular model $C^{min}$ of $C$ is isomorphic to $E^{min}$. Thus we will dispense with $C^{min}$ and write $E^{min}$ instead. The $S$-scheme ${\mathcal C}$ defined by an integral genus one equation $\phi:y^2+g(x,z)y=f(x,z)$ of degree $2$ is the scheme obtained by glueing $ \{ y^2+g(x,1)y = f(x,1) \} \subset {\mathbb A}_S^2 $ and $ \{ v^2+g(1,u)v = f(1,u) \} \subset {\mathbb A}_S^2 $ via $x=1/u$ and $y=x^2v$. It comes with a natural morphism ${\mathcal C} \to {\mathbb P}^1_S$ given on these affine pieces by $(x,y) \mapsto (x:1)$ and $(u,v) \mapsto (1:u)$. The $S$-scheme defined by an integral genus one equation $\phi$ of degree $n$, $n=1$ or $n\ge3$, is the subscheme ${\mathcal C} \subset {\mathbb P}^{m}_S$ defined by $\phi$, where $m=2$ when $n=1$, and $m=n-1$ when $n\ge3$. Now we give the key definition of this paper. \begin{Definition} \label{def:geometrically minimal} Let $\phi$ be an integral genus one equation of degree $n,\;n\ge 1$, defining a normal $S$-scheme ${\mathcal C}$. Assume moreover that ${\mathcal C}_K$ is smooth and ${\mathcal C}_K(K)\ne\emptyset$. Let $E^{min}$ be the minimal proper regular model of the Jacobian of ${\mathcal C}_K$. Then $\phi$ is said to be {\em geometrically minimal} if the minimal desingularisation $\widetilde{\mathcal C}\to{\mathcal C}$ satisfies $\widetilde{\mathcal C} \cong E^{min}$. \end{Definition} For the definitions of minimal proper regular models and minimal desingularisations see (\cite{Liubook}, \S 9.3). \section{Normality} \label{sec:normality} In this section we prove that an $S$-scheme defined by a minimal genus one equation of degree $n,\;n\le4,$ is normal, and hence is an $S$-model for its generis fiber. Let ${\mathcal C}$ be an $S$-scheme defined by an integral genus one equation of degree $n$ where ${\mathcal C}_K$ is smooth. It is known that the normality of ${\mathcal C}$ implies that there are only finitely many non-regular points on ${\mathcal C},$ and all these points are closed points in the special fiber. Moreover, if $n\le 4$, then ${\mathcal C}$ can be seen as a complete intersection. It follows that ${\mathcal C}$ is normal if and only if ${\mathcal C}$ is regular at the generic points of ${\mathcal C}_k.$ The latter statement is a direct consequence of Serre's criterion for normality, see (\cite{Liubook}, Corollary 8.2.24). If ${\mathcal C}_k$ is reduced, then ${\mathcal C}$ is normal, see (\cite{Liubook}, Lemma 4.1.18). \begin{Lemma} \label{lem:regularity} Let $(A,\mathfrak m)$ be a regular Noetherian local ring. \begin{myitemize} \item[(i)] Suppose that $f\in\mathfrak m\backslash\{0\}.$ Then $A/fA$ is regular if and only if $f\not\in\mathfrak m^2.$ \item[(ii)] Suppose that $I$ is a proper ideal of $A.$ Then $A/I$ is regular if and only if $I$ is generated by $r$ elements of $\mathfrak m$ which are linearly independent mod $\mathfrak m^2.$ \end{myitemize} \end{Lemma} \begin{Proof} See (\cite{Liubook}, Corollaries 4.2.12 and 4.2.15). \end{Proof} \begin{Lemma} \label{lem:S' normality implies S normality} Let $K'$ be a finite extension of $K$ with ring of integers $\mathcal{O}_{K'}$. Let ${\mathcal C}$ be an $S$-scheme. Set $S'=\operatorname{Spec}\mathcal{O}_{K'}$, and ${\mathcal C}'={\mathcal C}\times_SS'$. If ${\mathcal C}'$ is $S'$-normal, then ${\mathcal C}$ is $S$-normal. \end{Lemma} \begin{Proof} That ${\mathcal C}'$ is $S'$-normal means that $\mathcal{O}_{{\mathcal C}',x}$ is integrally closed in $\operatorname{Frac}(\mathcal{O}_{{\mathcal C}',x})$ for every $x\in{\mathcal C}'$. Now let $x\in{\mathcal C}$, and $\alpha\in\operatorname{Frac}(\mathcal{O}_{{\mathcal C},x})$ satisfy an integral relation for $\alpha$ over $\mathcal{O}_{{\mathcal C},x}$. The $S'$-normality of ${\mathcal C}'$ implies that $\alpha\in\mathcal{O}_{{\mathcal C}',x}$. Therefore, $\alpha\in\mathcal{O}_{{\mathcal C}',x}\cap\operatorname{Frac}({\mathcal{O}_{{\mathcal C},x}})=\mathcal{O}_{{\mathcal C},x}$. \end{Proof} If $f(x_1,\ldots,x_n)=\sum_{i=1}^m a_ix_1^{l_{1i}}\ldots x_n^{l_{ni}}\in{\mathcal{O}_K}[x_1,\ldots,x_n],$ then $\tilde{f}(x_1,\ldots,x_n)$ will denote its image in $k[x_1,\ldots,x_n].$ Moreover, $\nu(f)=\min \{\nu(a_i):1\le i\le m\}.$ Let ${\mathcal C}$ be an $S$-scheme defined by an integral genus one equation $\phi$ of degree $3$ where \begin{equation}\label{eq7} \phi: b y^3 + f_1(x,z)y^2+f_2(x,z) y+ f_3(x,z)=0, \end{equation} with $f_1(x,z)=b_1x+b_3z, f_2(x,z)=a_2x^2 + m x z +c_2 z^2$ and $ f_3(x,z)=ax^3+a_3x^2z+c_1z^2x+cz^3.$ If ${\mathcal C}_k$ contains an irreducible component of multiplicity-$m,\;m\ge 2,$ then we can assume without loss of generality that the defining equation of this multiplicity-$m$ component is $y=0.$ This means that $\min\{\nu(f_2),\nu(f_3)\}\ge1,\nu(f_1)=0$ when $m= 2,$ and $\min\{\nu(f_1),\nu(f_2),\nu(f_3)\}\ge1,\nu(b)=0$ when $m=3.$ \begin{Proposition} \label{prop:normal1,2,3} Let ${\mathcal C}$ be an $S$-scheme defined by an integral genus one equation $\phi$ of degree $n,\;n\le3.$ Assume that ${\mathcal C}_K$ is smooth. \begin{myitemize} \item[(i)] if ${\mathcal C}_k$ consists only of multiplicity-$1$ components, then ${\mathcal C}$ is normal. \end{myitemize} Now assume that ${\mathcal C}_k$ contains an irreducible component of multiplicity greater than $1$. \begin{myitemize} \item[(ii)] if $n=2$ and $\phi:y^2+g(x)y=f(x)$, then ${\mathcal C}$ is normal if and only if there exists $R(x)\in{\mathcal{O}_K}[x]$ such that $\nu(f(x)+g(x)R(x)-R(x)^2)=1$. \item[(iii)] if $n=3$, $\phi:F(x,y,z)=0$ is given as in equation (\ref{eq7}), and ${\mathcal C}_k$ contains a multiplicity-$m$ component $\Gamma:\{y=0\}$, $m\ge2$, then ${\mathcal C}$ is normal if and only if $\nu(f_3)=1$. \end{myitemize} \end{Proposition} \begin{Proof} $(i)$ Since ${\mathcal C}_k$ is reduced and ${\mathcal C}_K$ is smooth, the $S$-scheme ${\mathcal C}$ is normal. $(ii)$ This is Lemme 2 (c) of \cite{LiuModeles}. $(iii)$ The maximal ideal corresponding to the generic point $\xi$ of $\Gamma$ is $\mathfrak m_{\xi}=\langle t,y\rangle$. We have ${\mathcal C}$ is normal if and only if $F(x,y,z)\not\in\mathfrak m_{\xi}^2$, see Lemma \ref{lem:regularity} $(i)$. Since $\nu(f_2)\ge 1$, we have $y^3,y^2,f_2(x,z)y\in\mathfrak m_{\xi}^2.$ Therefore, $F(x,y,z)\not\in\mathfrak m_{\xi}^2$ if and only if $\nu(f_3)=1.$ \end{Proof} Now we study the normality of an $S$-scheme ${\mathcal C}$ defined by an integral genus one equation $\phi:F(x_1,x_2,x_3,x_4)=G(x_1,x_2,x_3,x_4)=0$ of degree $4$. We will assume that $\tilde{F}$ and $\tilde{G}$ are coprime. Moreover, we will assume that $\phi$ is not ${\mathcal{O}_K}$-equivalent to an equation whose reduction mod $t$ is given by $x_1^2=x_2^2=0$. These two assumptions are reasonable to make because of the following lemma which can be found in \S 2.5.1 of \cite{WomackThesis}. \begin{Lemma} \label{lem:quadruple line no rational points} \begin{myitemize} \item[(i)] If $\tilde{F}$ and $\tilde{G}$ have a common factor, then $\phi$ is not minimal. \item[(ii)] If ${\mathcal C}_k$ is defined by $x_1^2=x_2^2=0$, then either $\phi$ is not minimal, or ${\mathcal C}_K(K)=\emptyset$. \end{myitemize} \end{Lemma} We observe that (i) all the irreducible components of ${\mathcal C}_k$ are defined over the residue field of a finite unramified extension of $K$, (ii) the normality of ${\mathcal C}$ over the ring of integers of a finite extension of $K$ implies the normality of ${\mathcal C}$ over ${\mathcal{O}_K}$, see Lemma \ref{lem:S' normality implies S normality}, and (iii) the minimality of $\phi$ is stable under unramified base changes, see (\cite{FiStCr}, Theorem 3.6). Therefore, we will assume that $k$ is algebraically closed when we are finding criteria for the normality of ${\mathcal C}$, see Proposition \ref{prop:normal4}, and testing the normality of ${\mathcal C}$ when ${\mathcal C}$ is minimal, see Theorem \ref{thm:minimal+soluble imlies normal}. Since we will be interested in ${\mathcal C}$ when ${\mathcal C}_k$ contains a component of multiplicity-$m$, $m\ge 2$, we will write down all the possibilities for such a special fiber, up to ${\mathcal{O}_K}$-equivalence of $\phi$. Again $k$ will be algebraically closed for the remainder of this section. For a complete list for the forms of ${\mathcal C}_k$, which includes special fibers with only multiplicity-1 components, see \cite{parametrizationofquadrics}.\\ \begin{center} \begin{tabular}{|c|c|} \hline {${\mathcal C}_k$}&{Defining equations}\\ \hline conic + double line &$x_1x_3=x_2^2+x_1x_4=0$\\ \hline double conic& $x_1^2=x_2^2+x_3x_4=0$\\ \hline double line + two lines&$ x_1^2+x_2^2=x_1x_3+\mu x_2x_4=0,\;\mu\in \{0,1\}$\\ \hline triple line + line&$x_1x_2=x_1^2+x_2x_4=0$\\ \hline two double lines&$x_2^2=x_1x_3+\mu x_2x_4=0,\;\mu\in \{0,1\}$\\ \hline quadruple line&$x_1^2=x_2^2+ x_1x_3=0$\\ \hline \end{tabular} \end{center} Let the quadrics $F$ and $G$ be given by the following two polynomials respectively: {\setlength\arraycolsep{2pt} \begin{eqnarray} \label{eq8} a_1x_1^2&+&a_2x_1x_2+a_3x_1x_3+a_4x_1x_4+a_5x_2^2+a_6x_2x_3+a_7x_2x_4+a_8x_3^2+a_9x_3x_4+a_{10}x_4^2,\nonumber\\ b_1x_1^2&+&b_2x_1x_2+b_3x_1x_3+b_4x_1x_4+b_5x_2^2+b_6x_2x_3+b_7x_2x_4+b_8x_3^2+b_9x_3x_4+b_{10}x_4^2.\nonumber\\ \end{eqnarray}} \begin{Proposition} \label{prop:normal4} Let ${\mathcal C}$ be the $S$-scheme defined by the integral equation $\phi:F=G=0$, where $F$ and $G$ are given in (\ref{eq8}). Assume that ${\mathcal C}_K$ is smooth. \begin{myitemize} \item[(i)] If ${\mathcal C}_k$ contains a multiplicity-$1$ component $\Gamma,$ then ${\mathcal C}$ is normal at $\Gamma.$ \item[(ii)] If $\tilde{F}=x_1x_3$ and $\tilde{G}=x_2^2+x_1x_4,$ then ${\mathcal C}$ is normal if and only if $$\nu(x_4F(0,0,x_3,x_4)-x_3G(0,0,x_3,x_4))=1.$$ \item[(iii)] If $\tilde{F}=x_1^2$ and $\tilde{G}=x_2^2+x_3x_4.$ Then ${\mathcal C}$ is normal unless $F(0,x_2,x_3,x_4)\equiv\mu(x_2^2+x_3x_4)\mod t^2$, for some $\mu\in {\mathcal{O}_K}$. \item[(iv)] Assume that ${\mathcal C}_k$ contains a line $\Gamma:\{x_1=x_2=0\}$ of multiplicity-$m,\;m\ge2,$ with $\tilde{F}=q(x_1,x_2)$ and $\tilde{G}=x_1x_3+\mu x_2x_4+q'(x_1,x_2)$, $\mu\in\{0,1\}$. If $\nu(F(0,0,x_3,x_4))=1,$ then ${\mathcal C}$ is normal at $\Gamma$. \end{myitemize} \end{Proposition} \begin{Proof} $(i)$ Since ${\mathcal C}_k$ is reduced at the generic point $\xi$ of $\Gamma$, we see that ${\mathcal C}$ is normal at $\xi$. Now we use Lemma \ref{lem:regularity} $(ii)$ to study the normality of ${\mathcal C}$ at components of multiplicity greater than 1. The model ${\mathcal C}$ is normal if and only if $F,G\not\in\mathfrak m_{\xi}^2$, and $F,G$ are linearly independent mod $\mathfrak m_{\xi}^2,$ for every generic point $\xi$ of ${\mathcal C}_k.$ The linear independence condition is: For $\lambda_1,\lambda_2\in{\mathcal{O}_K}[x_1,\ldots,x_4]_{\xi},$ if $\lambda_1F+\lambda_2G\in\mathfrak m_{\xi}^2,$ then $\lambda_1,\lambda_2\in\mathfrak m_{\xi}.$ $(ii)$ Let $\xi$ be the generic point of the double line $\{x_1=x_2=0\}$ in ${\mathcal C}_k,$ then $\mathfrak m_{\xi}=\langle x_1,x_2,t\rangle.$ It is clear that $F,G\not\in\mathfrak m_{\xi}^2.$ If $\lambda_1F+\lambda_2G\in\mathfrak m_{\xi}^2,$ then the fact that $x_1$ and $t$ are linearly independent mod $\mathfrak m_{\xi}^2$ implies that $\lambda_1x_3+\lambda_2x_4\in\mathfrak m_{\xi},$ i.e., $\lambda_1\equiv \mu x_4 \mod\mathfrak m_{\xi}$ and $\lambda_2\equiv -\mu x_3\mod \mathfrak m_{\xi}$ for some $\mu\in\mathcal{O}_K.$ Thus ${\mathcal C}$ is normal if and only if $\nu(f)=1,$ where $$f=x_4(a_8x_3^2+a_9x_3x_4+a_{10}x_4^2)-x_3(b_8x_3^2+b_9x_3x_4+b_{10}x_4^2).$$ $(iii)$ Let $\mathfrak m_{\xi}=\langle x_1, x_2^2+x_3x_4,t\rangle$ be the maximal ideal corresponding to the generic point $\xi$ of the conic. We have $G\not\in\mathfrak m_{\xi}^2.$ If $a_5x_2^2+a_9x_3x_4=tu(x_2^2+x_3x_4),u\in\mathcal{O}_K,$ then $F\not\in\mathfrak m_{\xi}^2$ if and only if $\nu(a_6x_2x_3+a_7x_2x_4+a_8x_3^2+a_{10}x_4^2)=1,$ otherwise $F\not\in\mathfrak m_{\xi}^2$ if and only if $\nu(a_5x_2^2+a_6x_2x_3+a_7x_2x_4+a_8x_3^2+a_9x_3x_4+a_{10}x_4^2)=1.$ If $\lambda_1F+\lambda_2G\in\mathfrak m_{\xi}^2,$ then $\lambda_2\in\mathfrak m_{\xi}.$ The reason is $t$ and $x_2^2+x_3x_4$ are linearly independent mod $\mathfrak m_{\xi}^2.$ Then the condition we obtained from $F\not\in\mathfrak m_{\xi}^2$ implies that $\lambda_1\in\mathfrak m_{\xi}.$ $(iv)$ Assume that $\xi$ is the generic point of $\Gamma:\{x_1=x_2=0\}$. The ideal $\mathfrak m_{\xi}$ is given by $\langle x_1,x_2,t\rangle.$ Since $\tilde{F}=q(x_1,x_2)$ and $\nu(a_8x_3^2+a_9x_3x_4+a_{10}x_4^2)=1,$ we have $F\not\in\mathfrak m_{\xi}^2.$ Since $\tilde{G}=x_1x_3+\mu x_2x_4+q'(x_1,x_2)$, we have $G\not\in\mathfrak m_{\xi}^2$ because $x_1x_3\not\in\mathfrak m_{\xi}^2$. Assume that $\lambda_1F+\lambda_2G\in\mathfrak m_{\xi}^2.$ Since $x_1,x_2$ and $t$ are linearly independent mod $\mathfrak m_{\xi}^2,$ it follows that $\lambda_2\in\mathfrak m_{\xi}.$ Moreover, as $\nu(a_8x_3^2+a_9x_3x_4+a_{10}x_4^2)=1,$ we get $\lambda_1\in\mathfrak m_{\xi}$. \end{Proof} \begin{Theorem} \label{thm:minimal+soluble imlies normal} Let $\phi$ be an integral genus one equation of degree $n,\;n\le 4,$ defining an $S$-scheme ${\mathcal C}$. Assume that ${\mathcal C}_K$ is smooth and ${\mathcal C}_K(K)\ne\emptyset$. If $\phi$ is minimal, then ${\mathcal C}$ is normal. \end{Theorem} \begin{Proof} If ${\mathcal C}_k$ consists only of multiplicity-$1$ components, then ${\mathcal C}_k$ is reduced and hence ${\mathcal C}$ is normal. So, we only need to assume that $\phi$ is of degree $n,\;n\ge2,$ and ${\mathcal C}_k$ contains a component of multiplicity greater than $1.$ Furthermore, we assume that ${\mathcal C}$ is not normal, and hence $\phi$ does not satisfy the conditions included in Propositions \ref{prop:normal1,2,3} and \ref{prop:normal4}. Then we will prove that $\phi$ is not minimal by finding a genus one equation $K$-equivalent to $\phi$ whose discriminant has valuation less than $\Delta_{\phi}$. Recall that the discriminant varies by the 12th power of the determinant of the $K$-equivalence transformation, see \S \ref{sec:intro}. Let $n=2$ and $\phi:y^2+g(x)y=f(x)$. Since ${\mathcal C}_k$ is a double line, we can assume that $\tilde{g}=\tilde{f}=0$. Since $\phi$ is not normal, we have $\nu(f)\ge 2.$ The equation $\phi$ is not minimal because it is $K$-equivalent to the equation $y^2+\frac{1}{t}g(x)y=\frac{1}{t^2}f(x).$ Let $n=3$ and $\phi:F(x,y,z)=0$ as in Equation (\ref{eq7}). Now ${\mathcal C}_k$ contains a multiplicity-$m$ component $\Gamma:\{y=0\}$, $m\ge2$. Since ${\mathcal C}$ is not normal, we have $\nu(f_2),\nu(f_3)\ge1$ and $\nu(f_3)\ge 2$. Now $\phi$ is not minimal because it is $K$-equivalent to the equation $\frac{1}{t^2}F(x,ty,z)=0.$ Let $n=4$ and $\phi:F=G=0,$ where $F$ and $G$ are given as in (\ref{eq8}). We will go through the different cases of Proposition \ref{prop:normal4}. Assume that ${\mathcal C}_k:\{x_1x_3=x_2^2+x_1x_4=0\}$ and $\nu(x_4F(0,0,x_3,x_4)-x_3G(0,0,x_3,x_4))\ge2$. We use a matrix in $\operatorname{GL}_4({\mathcal{O}_K})$ to get rid of the $x_1^2,x_1x_2$ and $x_1x_4$-terms in $F$ and of the $x_1^2,x_1x_2$ and $x_1x_3$-terms in $G$. We notice that in the equation $$x_4F(0,0,x_3,x_4)-x_3G(0,0,x_3,x_4)=-b_8x_3^3+(a_8-b_9)x_3^2x_4+(a_9-b_{10})x_3x_4^2+a_{10}x_4^3,$$ we have $\min\{\nu(b_{8}),\nu(a_8-b_9),\nu(a_9-b_{10}),\nu(a_{10})\}\ge 2$. We apply the transformation $x_1\mapsto x_1-a_8x_3-a_9x_4,x_i\mapsto x_i,i=2,3,4,$ to get rid of the terms $a_8x_3^2$ and $a_9x_3x_4$. Thereafter, we obtain the genus one equation $\phi':F'=G'=0$, where {\setlength\arraycolsep{2pt} \begin{eqnarray} F'&=&x_1x_3+a_5x_2^2+a_6x_2x_3+a_7x_2x_4+a_{10}x_4^2,\nonumber\\ G'&=&x_1x_4+x_2^2+b_6x_2x_3+b_7x_2x_4+b_8x_3^2+(b_9-a_8)x_3x_4+(b_{10}-a_9)x_4^2.\nonumber \end{eqnarray}} We deduce that $\phi'$ is not minimal because it is $K$-equivalent to the equation $$\frac{1}{t^2}F'(t^2x_1,tx_2,x_3,x_4)=\frac{1}{t^2}G'(t^2x_1,tx_2,x_3,x_4)=0.$$ Assume that ${\mathcal C}_k:\{x_1^2=x_2^2+x_3x_4=0\},$ and that $\nu(F(0,x_2,x_3,x_4)-\mu(x_2^2+x_3x_4))\ge2$ for some $\mu\in{\mathcal{O}_K}$. Then $\phi$ is not minimal because it is $K$-equivalent to $$\frac{1}{t^2}(F(tx_1,x_2,x_3,x_4)-\mu G(tx_1,x_2,x_3,x_4))=G(tx_1,x_2,x_3,x_4)=0.$$ Now assume that ${\mathcal C}_k$ contains a line $\Gamma:\{x_1=x_2=0\}$ of multiplicity-$m,\;m\ge2,$ with $\tilde{F}=q(x_1,x_2)$ and $\tilde{G}=x_1x_3+\mu x_2x_4+q'(x_1,x_2),$ where $\mu\in\{0,1\}.$ Assume that $\nu(a_8x_3^2+a_9x_3x_4+a_{10}x_4^2)\ge2$. Then $\phi$ is not minimal because it is $K$-equivalent to $$\frac{1}{t^2}F(tx_1,tx_2,x_3,x_4)=\frac{1}{t}G(tx_1,tx_2,x_3,x_4)=0.$$ \end{Proof} \section{Criteria for minimality} \label{sec:criteria} We state the first main theorem of this paper. \begin{Theorem} \label{thm:minimality} Let $\phi$ be an integral genus one equation of degree $n,\;n\le4$. Assume moreover that $\phi$ defines a normal $S$-scheme ${\mathcal C}$, and that ${\mathcal C}_K$ is smooth with ${\mathcal C}_K(K)\ne\emptyset.$ Then $\phi$ is minimal, see Definition \ref{def:integral-minimal}, if and only if $\phi$ is geometrically minimal, see Definition \ref{def:geometrically minimal}. \end{Theorem} Theorem \ref{thm:minimality} is known for the case $n=1,$ see (\cite{Liubook}, \S 9.4) or \cite{Brian}. When $n=2$, Liu proved that if $\phi$ is minimal, then $\phi$ is geometrically minimal, see (\cite{LiuModeles}, Corollaire 5 (a)). We introduce a proof which works for $n$ when $n\le4.$ \subsection{Canonical sheaves of $S$-models} Let $E$ be an elliptic curve with minimal proper regular model $E^{min}$. If ${\mathcal C}$ is an $S$-model for $E,$ then the canonical sheaf $\omega_{{\mathcal C}/S}$ of ${\mathcal C}$ satisfies $\omega_{{\mathcal C}/S}|_E=\omega_{E/K}$, moreover the restriction of the canonical sheaf $\omega_{{\mathcal C}/S}$ on $E$ gives a canonical injection $H^0({\mathcal C},\omega_{{\mathcal C}/S})\hookrightarrow H^0(E,\omega_{E/K})$. We have $H^0({\mathcal C},\mathcal{O}_{{\mathcal C}})={\mathcal{O}_K}.$ In particular, if $\omega_{{\mathcal C}/S}=\omega\mathcal{O}_{{\mathcal C}},$ then $H^0({\mathcal C},\omega_{{\mathcal C}/S})=\omega{\mathcal{O}_K}.$ In addition, there exists an $\omega_0\in H^0(E^{min},\omega_{E^{min}/S})$ such that $\omega_{E^{min}/S}= \omega_0\mathcal{O}_{E^{min}},$ see (\cite{Brian}, Example 7.7). \begin{Lemma} \label{lem:H0Emin H0Cdelta} Let $E$ be an elliptic curve over $K$ with minimal proper regular model $E^{min}$. Let ${\mathcal C}$ be a normal $S$-scheme with ${\mathcal C}_K \cong E$. Let $\widetilde{\mathcal C}\to{\mathcal C}$ be the minimal desingularisation of ${\mathcal C}$. Then the following statements hold. \begin{myitemize} \item[(i)] $H^0(E^{min},\omega_{E^{min}/S})=H^0(\widetilde{\mathcal C},\omega_{\widetilde{\mathcal C}/S})\subseteq H^0({\mathcal C},\omega_{{\mathcal C}/S})$ as subgroups in $H^0(E,\omega_{E/K}).$ \item[(ii)] If $\widetilde{\mathcal C} \cong E^{min}$, then $H^0(E^{min},\omega_{E^{min}/S})=H^0({\mathcal C},\omega_{{\mathcal C}/S}).$ \end{myitemize} \end{Lemma} \begin{Proof} $(i)$ The equality holds because $\widetilde{{\mathcal C}}$ and $E^{min}$ are two regular $S$-models for $E$, see (\cite{Liubook}, Corollary 9.2.25 (b)). The inequality holds because ${\mathcal C}$ is obtained from $\widetilde{\mathcal C}$ as a contraction of a finite number of irreducible components, see (\cite{Liubook}, Lemma 9.2.17 (a)). $(ii)$ Since $\widetilde{\mathcal C} \cong E^{min}$, we have a contraction morphism $f:E^{min}\to {\mathcal C}$. Therefore, $f_*\omega_{E^{min}/S}=\omega_{{\mathcal C}/S}$, see (\cite{Liubook}, Corollary 9.4.18 (b)). \end{Proof} \begin{Proposition} \label{prop:canoniacal sheaf of degree n model} Let ${\mathcal C}$ be an $S$-scheme defined by an integral genus one equation $\phi$ of degree $n$. Assume that ${\mathcal C}_K$ is smooth. Then $\omega_{{\mathcal C}/S}=\omega_{\phi} \mathcal{O}_{{\mathcal C}},$ where $\omega_{\phi}\in H^0({\mathcal C}_K,\omega_{{\mathcal C}_K/K})$ is \begin{myitemize} \item[(i)] if $n=1$: $\omega_{\phi}= du/(2v+a_1u+a_3),\textrm{ where }u=x/z,v=y/z\in K({\mathcal C}),$ \item[(ii)] if $n=2$: $\omega_{\phi}= dx/(2y+g(x)),$ \item[(iii)] if $n=3$: $\omega_{\phi}=du/(\partial{F}/\partial{v}),\textrm{ where } u=x/z, v=y/z\in K({\mathcal C}), $ \item[(iv)] if $n=4$: $\omega_{\phi}= d u_2/ (\frac{\partial{F_1}}{\partial{u_4}}\frac{\partial{F_2}}{\partial{u_3}}-\frac{\partial{F_1}}{\partial{u_3}}\frac{\partial{F_2}}{\partial{u_4}}), \textrm{ where }u_i=x_i/x_1 \in K({\mathcal C}),\;i=2,3,4.$ \end{myitemize} \end{Proposition} \begin{Proof} This is a direct consequence of (\cite{Liubook}, Corollary 6.4.14). \end{Proof} \begin{Lemma} \label{lem:rel between delta.omega} Let $\phi_1,\phi_2$ be two $K$-equivalent integral genus one equations of degree $n,\;n\le4$. Assume that $\phi_i$ defines an $S$-scheme ${\mathcal C}_i$ with $C_i:=({\mathcal C}_i)_K$ being smooth. If $\omega_{{\mathcal C}_i/S}=\omega_{\phi_i}\mathcal{O}_{{\mathcal C}_i}$, then $$\Delta_{\phi_1}\omega_{\phi_1}^{\otimes 12}=\lambda\Delta_{\phi_2} \omega_{\phi_2}^{\otimes 12}\in H^0(C_1,\omega_{C_1/K})^{\otimes 12}=H^0(C_2,\omega_{C_2/K})^{\otimes12},\textrm{ where }\lambda\in\mathcal{O}^*_K.$$ \end{Lemma} \begin{Proof} Assume that $\phi_1=g.\phi_2$ where $g$ is a transformation defining the $K$-equivalence. The transformation $g$ defines an isomorphism $\gamma:C_1 \cong C_2$ which satisfies $\gamma^*\omega_{\phi_2}=(\det g) \omega_{\phi_1},$ see (\cite{FiInvariants}, Proposition 5.19). Hence, $\omega_{\phi_2}=\alpha(\det g) \omega_{\phi_1}$ as elements in $H^0(C_i,\omega_{C_i/K}),$ where $\alpha\in\mathcal{O}^*_K.$ Recall that $\Delta_{\phi_1}=(\det g)^{12}. \Delta_{\phi_2}.$ It follows that $\Delta_{\phi_1}\omega_{\phi_1}^{\otimes 12}=\alpha^{12}\Delta_{\phi_2} \omega_{\phi_2}^{\otimes 12}.$ \end{Proof} If $\phi_1$ is minimal, then we call the integer $m$ such that $\omega_{\phi_2}=ut^{-m}\omega_{\phi_1},u\in\mathcal{O}^*_K,$ the {\em level} of $\phi_2,$ and denote it by $\operatorname{level}(\phi_2).$ The above corollary implies that the level of an integral genus one equation of degree $n$ does not depend on the choice of the minimal genus one equation $\phi_1.$ Notice that $\nu(\Delta_{\phi_2})=\nu(\Delta_{\phi_1})+12 \operatorname{level}(\phi_2).$ \begin{Lemma} \label{lem:H0 subset H0} Keep the hypothesis of Lemma \ref{lem:rel between delta.omega}. Then we have $H^0({\mathcal C}_1,\omega_{{\mathcal C}_1/S})\subseteq H^0({\mathcal C}_2,\omega_{{\mathcal C}_2/S})$ as sub-${\mathcal{O}_K}$-modules of $H^0(C_i,\omega_{C_i/K})$ if and only if $\nu(\Delta_{\phi_1})\le\nu(\Delta_{\phi_2}).$ Moreover, the equality of the two submodules holds if and only if $\phi_1$ and $\phi_2$ have the same level. \end{Lemma} \begin{Proof} The assumption $H^0({\mathcal C}_1,\omega_{{\mathcal C}_1/S})\subseteq H^0({\mathcal C}_2,\omega_{{\mathcal C}_2/S})$ is equivalent to $\omega_{\phi_1}\in\omega_{\phi_2}{\mathcal{O}_K}.$ Thus Lemma \ref{lem:rel between delta.omega} asserts that $\Delta_{\phi_2}\in\Delta_{\phi_1}{\mathcal{O}_K},$ i.e., $\nu(\Delta_{\phi_1}))\le\nu(\Delta_{\phi_2}).$ The equality of the sub-${\mathcal{O}_K}$-modules $H^0({\mathcal C}_1,\omega_{{\mathcal C}_1/S})= H^0({\mathcal C}_2,\omega_{{\mathcal C}_2/S})$ means that $\omega_{\phi_1}{\mathcal{O}_K}=\omega_{\phi_2}{\mathcal{O}_K}$ as ${\mathcal{O}_K}$-modules, i.e., $\omega_{\phi_1}\in\omega_{\phi_2}\mathcal{O}^*_K$. Hence, $\phi_1$ and $\phi_2$ have the same level by Lemma \ref{lem:rel between delta.omega}. \end{Proof} \subsection{Constructing genus one equations} \label{sec:constructing minimal degree n models} Let $E$ be an elliptic curve over $K$ with minimal proper regular model $E^{min}$. Let $P\in E(K)$. We will denote the Zariski closure of $\{P\}$ in $E^{min}$ by $\overline{\{P\}}$. When $n=1,$ set $D_1=3.\overline{\{P\}}$ where $P\in E(K).$ When $n\ge2,$ let $\sum(P_i)\in\operatorname{Div}(E)$ be a $K$-rational divisor on $E$ of degree $n.$ Assume moreover that $\overline{\{P_i\}}\cap E^{min}_k$ is contained in one and only one irreducible component of $E^{min}_k.$ Consider the divisor $D_n$ on $E^{min}$, and the $S$-model ${\mathcal C}_n$ for $E$ given by $$D_n=\sum\overline{\{P_i\}},\textrm{ and }{\mathcal C}_n:=\operatorname{Proj}(\bigoplus_{m=0}^{\infty}H^0(E^{min},\mathcal{O}_{E^{min}}(mD_n))).$$ There is a canonical morphism $u_n:E^{min}\rightarrow {\mathcal C}_n$ contracting all the irreducible components of $E^{min}_k$ except the ones having nonempty intersection with $D_n$, see Theorem \ref{th:contraction}. \begin{Lemma} \label{lem:free module} Let $D_n, \;n\in\{1,2,3,4\},$ be as above. Then $H^0(E^{min},\mathcal{O}_{E^{min}}(mD_n)),m\ge1,$ is a free ${\mathcal{O}_K}$-module of rank $3m$ if $n=1,$ and of rank $mn$ if $n\ge2.$ \end{Lemma} \begin{Proof} It is known that $H^0(E^{min},\mathcal{O}_{E^{min}}(mD_n))\otimes_{{\mathcal{O}_K}}K \cong H^0(C,\mathcal{O}_C(mD_n|_C))$, see for example (\cite{Liubook}, Corollary 5.2.27). By virtue of the Riemann-Roch Theorem, $H^0(C,\mathcal{O}_C(mD_n|_C))$ is a $3m$-dimensional $K$-vector space when $n=1,$ and an $mn$-dimensional $K$-vector space when $n\ge2.$ Since $\mathcal{O}_{E^{min}}(mD_n)$ is an invertible sheaf on $E^{min}$, it follows that $H^0(E^{min},\mathcal{O}_{E^{min}}(mD_n))$ is a flat ${\mathcal{O}_K}$-module. Therefore, $H^0(E^{min},\mathcal{O}_{E^{min}}(mD_n))$ is a finitely generated flat module over the local ring ${\mathcal{O}_K}$, hence it is free. \end{Proof} \begin{Theorem} \label{thm:C_n is minimal} Let ${\mathcal C}_n$ and $D_n,\;n\le4,$ be as above. Then there exists an integral genus one equation $\phi_n$ of degree $n$ defining ${\mathcal C}_n$. \end{Theorem} \begin{Proof} For $ n=1,$ see (\cite{Liubook}, \S 9.4). For $n\ge2$, we pick a basis $\{x_1,\ldots,x_n\}$ of $H^0(E^{min},\mathcal{O}_{E^{min}}(D_n))$. Consider the morphism $\lambda_n:E^{min}\longrightarrow {\mathbb P}_S^{n-1}$ associated to the basis $\{x_1,\ldots,x_n\}$. For $n=2$, put $x_1=x$ and $x_2=1$. We have ${\mathbb P}_S^1=\operatorname{Spec}{\mathcal{O}_K}[x]\cup\operatorname{Spec}{\mathcal{O}_K}[1/x]$. Let $U=\lambda_2^{-1}(\operatorname{Spec}{\mathcal{O}_K}[x])$ and $V=\lambda_2^{-1}(\operatorname{Spec}{\mathcal{O}_K}[1/x])$. We have ${\mathcal C}_2=U\cup V.$ Taking the integral closure of ${\mathcal{O}_K}[x]$ in $K({\mathcal C}_2)$, we have $$\mathcal{O}_{{\mathcal C}_2}(U)={\mathcal{O}_K}[x]\oplus y{\mathcal{O}_K}[x], \textrm{ for some }y\in\mathcal{O}_{{\mathcal C}_2}(U),$$ moreover there exist $g(x),f(x)\in{\mathcal{O}_K}[x]$ such that $\deg g\le 2,\;\deg f \le 4$ and $y^2+g(x)y=f(x)$, see (\cite{LiuModeles}, Lemme 1). Following the same argument for $\mathcal{O}_{{\mathcal C}_2}(V)$, we deduce that ${\mathcal C}_2$ is the union of the two affine open schemes \begin{eqnarray} U=\operatorname{Spec}{\mathcal{O}_K}[x,y]/(y^2+g(x)y-f(x)),\textrm{ }V=\operatorname{Spec}{\mathcal{O}_K}[w,z]/(z^2+w^2 g(1/w)z-w^4f(1/w)),\nonumber \end{eqnarray} where $w=1/x,\;z=y/x^2.$ Hence, we are done when $n=2$. Now let $n=3,4$. Let $Z_n$ be the closed subset $\lambda_n(E^{min})\subset{\mathbb P}^{n-1}_S$ endowed with the reduced scheme structure. According to the description of the contraction morphism included in the proof of (\cite{Liubook}, Proposition 8.3.30), the morphism $\lambda_n:E^{min}\to Z_n\subseteq {\mathbb P}_S^{n-1}$ factors into $u_n:E^{min}\to {\mathcal C}_n$ followed by $v_n:{\mathcal C}_n\to Z_n,$ where $v_n$ is the normalisation morphism. It is understood that $v_n$ is a finite morphism, hence for an irreducible component $\Gamma$ of $E^{min}_k$, $\lambda_n(\Gamma)$ is a point if and only if $u_n(\Gamma)$ is a point. In other words, the special fibers of ${\mathcal C}_n$ and $Z_n$ have the same number of irreducible components. We are going to show that $Z_n$ is defined by an integral genus one equation of degree $n$. Then we show that ${\mathcal C}_n \cong Z_n$. When $n=3$, the free ${\mathcal{O}_K}$-module $H^0(E^{min},\mathcal{O}_{E^{min}}(3D_3))$ is of rank-$9$, see Lemma \ref{lem:free module}, but it contains the $10$ elements $x_1^3, x_2^3,x_3^2, x_1^2x_2, x_1^2x_3, x_2^2x_1, x_2^2x_3, x_3^2x_1, x_3^2x_2, x_1x_2x_3$. It follows that there are $a_i\in{\mathcal{O}_K}$ such that $$F:=a_1x_1^3+a_2x_2^3+a_3x_3^3+a_4x_1^2x_2+a_5x_1^2x_3+a_6x_2^2x_1+a_7 x_2^2x_3+a_8 x_3^2x_1+a_9 x_3^2x_2+a_{10}x_1x_2x_3=0.$$ Rescaling $x_1,\;x_2$ and $x_3$, we can assume that there is at least one $a_i\in\mathcal{O}^*_K$. Now $Z_3$ is contained in $\operatorname{Proj}{\mathcal{O}_K}[x_1,x_2,x_3]/(F)$. When $n=4$, we consider the $10$ elements $x_1^2,x_1x_2,x_1x_3,x_1x_4,x_2^2,x_2x_3,x_2x_4,x_3^2,x_3x_4,x_4^2$ in the rank-$8$ free ${\mathcal{O}_K}$-module $H^0(E^{min},\mathcal{O}_{E^{min}}(2D_4))$. They satisfy two quadrics $Q$ and $R$ with coefficients in ${\mathcal{O}_K}$. Moreover, by rescaling $x_1,x_2,x_3$ and $x_4$ we can assume that both $Q$ and $R$ have at least one coefficient in $\mathcal{O}_K^*$. Now $Z_4$ is contained in the intersection of $Q$ and $R$. We want to show that $Z_n= \operatorname{Proj}{\mathcal{O}_K}[x_1,\ldots,x_n]/I_n$, where $I_3=(F)$ and $I_4=(Q,R)$. Recall that both schemes are of dimension 2. Since $Z_n\subseteq\operatorname{Proj}{\mathcal{O}_K}[x_1,\ldots,x_n]/I_n$, we have $\operatorname{Proj}{\mathcal{O}_K}[x_1,\ldots,x_n]/I_n=Z_n\cup Z_n'$, for some closed subscheme $Z_n'\subset{\mathbb P}^{n-1}_S,$ where $Z_n'\ne \operatorname{Proj}{\mathcal{O}_K}[x_1,\ldots,x_n]/I_n$. Since $\operatorname{Proj}({\mathcal{O}_K}[x_1,\ldots,x_n]/I_n\otimes K)$ is irreducible, $\operatorname{Proj}{\mathcal{O}_K}[x_1,\ldots,x_n]/I_n$ is irreducible itself. It follows from the definition of irreducibility that $Z_n'=\emptyset$, and the closed subscheme $Z_n$ is $\operatorname{Proj}{\mathcal{O}_K}[x_1,\ldots,x_n]/I_n$. Now both ${\mathcal C}_n$ and $Z_n$, $n=3,4,$ have dimension $2$, their generic fibers are isomorphic, and their special fibers have the same number of irreducible components. By virtue of (\cite{Liubook}, Remark 8.3.25), $v_n:{\mathcal C}_n\to Z_n$ is an isomorphism. \end{Proof} \begin{Remark} \label{rem1} Let $n\ge2.$ Let $\phi$ be an integral genus one equation of degree $n$ defined by a hyperplane section $H.$ Assume moreover that $\phi$ defines a morphism $E\to{\mathbb P}^{n-1}_K$. Then we can choose $D_n=(n-1)\overline{\{P\}}+\overline{\{Q\}}$ where $P,Q\in E(K)$ are such that $(n-1)P+Q\sim H.$ It follows that $\phi_n$ is $K$-equivalent to $\phi$. \end{Remark} \subsection{Proof of Theorem \ref{thm:minimality}} \begin{ProofOf}{Theorem \ref{thm:minimality}} We first assume that $\phi$ is geometrically minimal, i.e., that $\widetilde{{\mathcal C}} \cong E^{min}$. Thus we have a contraction morphism $f : E^{min}\rightarrow {\mathcal C}.$ We claim that if ${\mathcal C}'$ is a normal $S$-scheme defined by an integral genus one equation $\phi'$ which is $K$-equivalent to $\phi$, then $H^0({\mathcal C},\omega_{{\mathcal C}/S})\subseteq H^0({\mathcal C}',\omega_{{\mathcal C}'/S})$, and hence $\nu(\Delta_{\phi})\le\nu(\Delta_{\phi'}),$ see Lemma \ref{lem:H0 subset H0}. Therefore, $\phi$ is minimal. Lemma \ref{lem:H0Emin H0Cdelta} $(i)$ shows that $H^0(E^{min},\omega_{E^{min}/S})\subseteq H^0({\mathcal C}',\omega_{{\mathcal C}'/S}).$ The fact that ${\mathcal C}$ is obtained from $E^{min}$ by contraction implies that $H^0({\mathcal C},\omega_{{\mathcal C}/S}) = H^0(E^{min},\omega_{E^{min}/S})$, see Lemma \ref{lem:H0Emin H0Cdelta} $(ii)$. Thus the claim is proved. Before we proceed with the proof of the second part of the theorem we need the following lemma. \begin{Lemma} \label{lem:H0 Emin and H0 C when min} Let $\phi$ be a genus one equation of degree $n$ defining a normal $S$-scheme ${\mathcal C}$ with ${\mathcal C}_K$ being smooth and ${\mathcal C}_K(K)\ne\emptyset.$ Let $E^{min}$ be the minimal proper regular model of the Jacobian of ${\mathcal C}_K$. Then $H^0(E^{min},\omega_{E^{min}/S})= H^0({\mathcal C},\omega_{{\mathcal C}/S})$ as sub-${\mathcal{O}_K}$-modules of $H^0({\mathcal C}_K,\omega_{{\mathcal C}_K/K})$ if and only if $\phi$ is minimal. \end{Lemma} \begin{Proof} Assume that $H^0(E^{min},\omega_{E^{min}/S})=H^0({\mathcal C},\omega_{{\mathcal C}/S}).$ Let ${\mathcal C}'$ be a normal $S$-scheme defined by a genus one equation $K$-equivalent to $\phi$. By virtue of Lemma \ref{lem:H0Emin H0Cdelta} $(i)$, we have $H^0({\mathcal C},\omega_{{\mathcal C}/S})\subseteq H^0({\mathcal C}',\omega_{{\mathcal C}'/S})$. Therefore, $\phi$ is minimal by Lemma \ref{lem:H0 subset H0}. Now assume that $\phi$ is minimal. According to Theorem \ref{thm:C_n is minimal} and Remark \ref{rem1}, there exists an integral genus one equation $\phi'$ which is $K$-equivalent to $\phi$. Moreover, $\phi'$ is geometrically minimal because the minimal desingularisation of the $S$-scheme ${\mathcal C}'$ defined by $\phi'$ is isomorphic to $E^{min}.$ Therefore, $H^0(E^{min},\omega_{E^{min}/S})=H^0({\mathcal C}',\omega_{{\mathcal C}'/S})$ by Lemma \ref{lem:H0Emin H0Cdelta} $(ii)$. Moreover, $\phi'$ is minimal by the first part of Theorem \ref{thm:minimality}. Since the genus one equations $\phi$ and $\phi'$ are both minimal, in particular they have the same level, Lemma \ref{lem:H0 subset H0} implies that $H^0({\mathcal C}',\omega_{{\mathcal C}'/S})=H^0({\mathcal C},\omega_{{\mathcal C}/S})$, and we are done. \end{Proof} Now we conclude the proof of Theorem \ref{thm:minimality}. Assume that $\phi$ is minimal. We assume on the contrary that $\widetilde{\mathcal C}\not \cong E^{min},$ and therefore ${\mathcal C}_k$ contains an exceptional divisor $\Gamma.$ By (\cite{Liubook}, Proposition 9.3.10), we have $\deg \omega_{\widetilde{\mathcal C}/S}|_{\Gamma}<0$. It follows that $H^0(\Gamma,\omega_{\widetilde{\mathcal C}/S}|_{\Gamma})=0,$ therefore $\omega_{\widetilde{\mathcal C}/S}$ is not generated by its global sections on $\Gamma$. But we have $$\omega_{\widetilde{\mathcal C}/S}|_{\Gamma}=\omega_{{\mathcal C}/S}|_{\Gamma}=\omega_{\phi}\mathcal{O}_{{\mathcal C}}|_{\Gamma},$$ where $\omega_{\phi}$ is given as in Proposition \ref{prop:canoniacal sheaf of degree n model}. The global sections of $\omega_{\widetilde{\mathcal C}/S}$ are $$H^0(\widetilde{\mathcal C},\omega_{\widetilde{\mathcal C}/S})=H^0(E^{min},\omega_{E^{min}/S})=H^0({\mathcal C},\omega_{{\mathcal C}/S})=\omega_{\phi}{\mathcal{O}_K},$$ where the second equality is justified by ${\mathcal C}$ being minimal, see Lemma \ref{lem:H0 Emin and H0 C when min}. Therefore, $\omega_{\widetilde{\mathcal C}/S}$ is generated by its global sections at every $x\in\Gamma,$ which is a contradiction. \end{ProofOf} \begin{Corollary} \label{cor:critieria for minimality} Let $\phi,{\mathcal C}$ and $\widetilde{\mathcal C}$ be as in Theorem \ref{thm:minimality}, and $\omega_{\phi}$ as in Proposition \ref{prop:canoniacal sheaf of degree n model}. Then the following are equivalent.\\ \begin{tabular}{p{7cm}p{8cm}} (i) $\phi$ is minimal & (ii) $\phi$ is geometrically minimal \\ (iii) $\omega_{\widetilde{\mathcal C}/S}=\omega_{\phi}\mathcal{O}_{\widetilde{\mathcal C}}$ & (iv) $H^0(\widetilde{\mathcal C},\omega_{\widetilde{\mathcal C}/S})=H^0({\mathcal C},\omega_{{\mathcal C}/S})=\omega_{\phi}{\mathcal{O}_K}.$ \end{tabular} \end{Corollary} \section{Existence of global models} \label{sec:global models} In order to prove the global result included in Theorem \ref{thm1}, we need the following local result. \begin{Lemma} \label{lem:ord delta} Let $\phi$ be a minimal genus one equation of degree $n$ defining a normal $S$-scheme ${\mathcal C}$ such that ${\mathcal C}_K$ is smooth and ${\mathcal C}_K(K)\ne\emptyset.$ Let $E$ be the Jacobian of ${\mathcal C}_K$ with minimal discriminant $\Delta$. Then $\nu(\Delta_{\phi})=\nu(\Delta)$. \end{Lemma} \begin{Proof} Let ${\mathcal E}$ and $E^{min}$ be the minimal Weierstrass model and minimal proper regular model of $E$ respectively. We identify ${\mathcal C}_K$ and $E$ via an isomorphism $\gamma:{\mathcal C}_K \cong _K E$. For explicit formulae for $\gamma$, see (\cite{FiInvariants}, \S 6) or (\cite{FiStCr}, \S 2). Let $\omega_{\phi}$ and $\omega$ be the generators of the canonical sheaves $\omega_{{\mathcal C}/S}$ and $\omega_{{\mathcal E}/S}$ given as in proposition \ref{prop:canoniacal sheaf of degree n model}. Now according to Lemma \ref{lem:H0 Emin and H0 C when min}, we have $H^0(E^{min},\omega_{E^{min}/S})=H^0({\mathcal E},\omega_{{\mathcal E}/S})=\omega{\mathcal{O}_K}$, and $H^0(E^{min},\omega_{E^{min}/S})=H^0({\mathcal C},\omega_{{\mathcal C}/S})=\omega_{\phi}{\mathcal{O}_K}$ as sub-${\mathcal{O}_K}$-modules of $H^0(E,\omega_{E/K})$. Therefore, $\omega_{\phi}=\lambda\omega$ for some $\lambda\in\mathcal{O}_K^*.$ According to Lemma \ref{lem:rel between delta.omega}, we have $\Delta_{\phi}.\omega_{\phi}^{\otimes12}=\Delta.\omega^{\otimes12}$ up to a unit. Hence, $\Delta_{\phi}=\Delta$ up to a unit in $\mathcal{O}_K^*$. In other words, $\nu(\Delta_{\phi})=\nu(\Delta).$ \end{Proof} \begin{Corollary} \label{cor:delta and delta min} Let $\phi$ be an integral genus one equation defining a normal $S$-scheme ${\mathcal C}$ such that ${\mathcal C}_K$ is smooth and ${\mathcal C}_K(K)\ne\emptyset.$ Let $\widetilde{\mathcal C}\to{\mathcal C}$ be the minimal desingularisation of ${\mathcal C}$. Let $E$ be the Jacobian of ${\mathcal C}_K$ with minimal discriminant $\Delta$. Then $$\nu(\Delta_{\phi})=\nu(\Delta)+12\operatorname{length}_{{\mathcal{O}_K}}(H^0({\mathcal C},\omega_{{\mathcal C}/S})/H^0(\widetilde{\mathcal C},\omega_{\widetilde{\mathcal C}/S})),$$ where $\operatorname{length}_{{\mathcal{O}_K}}$ denotes the length of an ${\mathcal{O}_K}$-module. \end{Corollary} \begin{Proof} Theorem \ref{thm:C_n is minimal} and Remark \ref{rem1} admit the existence of an integral geometrically minimal genus one equation $\phi'$ $K$-equivalent to $\phi$. Moreover, Theorem \ref{thm:minimality} implies that $\phi'$ is minimal. We have $\nu(\Delta_{\phi'})=\nu(\Delta)$ by Lemma \ref{lem:ord delta}. We only need to show that $\nu(\Delta_{\phi})=\nu(\Delta_{\phi'})+12\operatorname{length}_{{\mathcal{O}_K}}(H^0({\mathcal C},\omega_{{\mathcal C}/S})/H^0(\widetilde{\mathcal C},\omega_{\widetilde{\mathcal C}/S}))$. The latter follows immediately from the fact that $\Delta_{\phi}.\omega_{\phi}^{\otimes 12}=\Delta_{\phi'}.\omega_{\phi'}^{\otimes 12}$ up to a unit, see Lemma \ref{lem:rel between delta.omega}, and $H^0({\mathcal C}',\omega_{{\mathcal C}'/S})=H^0(\widetilde{\mathcal C},\omega_{\widetilde{\mathcal C}/S})\subseteq H^0({\mathcal C},\omega_{{\mathcal C}/S})=\omega_{\phi}\mathcal{O}_{{\mathcal C}}$, where ${\mathcal C}'$ is the normal $S$-scheme defined by $\phi'$. \end{Proof} We observe that $\operatorname{length}_{{\mathcal{O}_K}}(H^0({\mathcal C},\omega_{{\mathcal C}/S})/H^0(\widetilde{\mathcal C},\omega_{\widetilde{\mathcal C}/S}))$ in the above corollary is an interpretation for $\operatorname{level}(\phi)$ mentioned after Lemma \ref{lem:rel between delta.omega}. Now we conclude with the proof of Theorem \ref{thm1}. \begin{ProofOf}{Theorem \ref{thm1}} We write $\mathcal{O}_{\nu}$ for the ring of integers of the completion $F_{\nu}$. Set $S_{\nu}=\operatorname{Spec}\mathcal{O}_{\nu}$. By virtue of Theorem \ref{thm:C_n is minimal} and Remark \ref{rem1}, $\phi$ is $F_{\nu}$-equivalent to a geometrically minimal $\mathcal{O}_{\nu}$-integral genus one equation $\phi_{\nu}$. Moreover, $\phi_{\nu}$ is minimal, see Theorem \ref{thm:minimality}. According to Lemma \ref{lem:ord delta}, $\nu(\Delta_{\phi_{\nu}})=\nu(\Delta)$ for every non-archimedean place $\nu$. Since $F$ is of class number $1$, we can use strong approximation to find an $\mathcal{O}_F$-integral genus one equation $\phi'$ which is $F$-equivalent to $\phi$ and $\nu(\Delta_{\phi'})=\nu(\Delta_{\phi_{\nu}})=\nu(\Delta)$, for every non-archimedean place $\nu$. See \cite{Fiminimise} for details when $n=2,3$. The case $n=4$ is similar. Since the discriminant is of weight 12, we have $\Delta_{\phi'}=\lambda^{12}\Delta$ for some $\lambda\in\mathcal{O}_F^*.$ By rescaling the coefficients of the defining polynomials of $\phi'$, we can assume that $\lambda=1.$ \end{ProofOf} \hskip-18pt\emph{\bf{Acknowledgements.}} This paper is based on the author's Ph.D. thesis \cite{SadekThesis} at Cambridge University. The author would like to thank his supervisor Tom Fisher for guidance and many useful comments on the manuscript. \bibliographystyle{plain} \footnotesize
\section{Introduction} Simulations that assume simple, non-interacting dark matter successfully reproduce cosmic large-scale structure \citep[see e.g.][]{sfw06}. However, they also predict power-law cusps of dark density at the centres of all virialised structures \citep[e.g.][]{dubinski1991,nfw1996,moore1998,diemand2004,diemand2005}. Dynamical and other circumstantial evidence for many galaxy types imply cores of nearly uniform dark density \citep[e.g][]{flores1994,moore1994,burkert1995,weldrake2003, gentile2004,simon2005,deblok2005,kuzio2006,weijmans2008}. It remains debatable whether the present-day cores are innate, or whether a violent, primordial form of baryonic ``feedback'' somehow conspired to erase (or exacerbate) all cusps \citep[e.g.][]{blumenthal1986,gnedin2002,gnedin2004, romano-diaz2008,jardel2008}. The observational situation in elliptical galaxies is less clear than for other morphological types, because of the relative difficulty of observing suitable kinematic tracers at a range of radii. These data may suffer degeneracies between mass profiles and the orbital anisotropies of kinematic tracers \citep{binney82}. X-ray emission (if detected) is informative if hydrostasis is a fair assumption. Observations of some X-ray emitting elliptical galaxies favour the presence of massive dark halos \citep[e.g.][]{awaki1994,loewenstein1999,mathews2003,humphrey2006}. For some controversial galaxies, dark matter appears dynamically absent, perhaps more diffuse than in standard cosmogonies, or possibly disguised by stellar orbital anisotropies \citep[e.g.][]{mendez2001,romanowsky2003,dekel2005,delorenzi2008b}. In other cases the inference of dark matter is clearer \citep[e.g][]{osullivan2004,delorenzi2008a}. Gravitational lens models suggest steep, isothermal slopes around the observed radii. However, inner regions show moderate dark matter densities \citep[e.g.][]{gerhard2001,ferreras2005,cappellari2006,thomas2007,weijmans2009}, despite the cuspy predictions of simple collisionless dark matter cosmogonies, and the expectation of adiabatic contraction due to the infall of the baryons towards the centre of the halo \citep[e.g.][]{blumenthal1986}. Before N-body simulations became popular and highly resolved, flat-cored profiles like the ``pseudo-isothermal sphere'' were by default assumed in studies of galaxies and galaxy clusters. Observations of galaxy-scale dark cores now motivate the examination of alternative dark matter theories where the physics naturally cause cores, regardless of any baryonic ``feedback'' \citep[e.g.][]{spergel2000,peebles2000,arbey2003,ahn2005,saxton2008}. \citet{nunez2006} and \citet{zavala2006} compared the scaling relations of disc galaxies using either a standard dark matter profile \citep[NFW,][]{nfw1996}, or a polytrope -- intended to represent collisionless dark matter subject to Tsallis thermostatistics \citep{tsallis1988} -- and found the latter to explain better the observations. \citet{boehmer2007} also found good fits to the rotation curves of low surface brightness galaxies and dwarf galaxies with a polytropic dark matter halo (which they justify via the formation of a Bose-Einstein condensate). Furthermore, polytropic dark matter halos can equally represent other interesting scenarios, such as self-interacting dark matter, where the core has formed due to the support of dark pressure in the dark field domain. Polytropic models are naturally consistent, self-gravitating distributions of classical mass, and may be seen as analogues to adiabatic fluids. Some early simulations with numerical approximations to dark hydrodynamics \citep{moore2000,yoshida2000b,burkert2000,kochanek2000} entailed (dark) thermal conduction leading to gravothermal catastrophe, and hence isothermal cusps that were sharper than those of collisionless simulations. However it was eventually shown \citep{balberg2002a,balberg2002b,ahn2005} that the simulations were unwittingly (and inappropriately) initiated in near-collapse conditions. It was shown that conductivity is only significant for SIDM of medium interactivity. Weaker SIDM has infrequent collisions and conduction is slow. Strongly self-interactive dark matter has vanishing conductivity (as it returns to an adiabatic limit) which defers collapse well beyond the age of the universe. In this paper we extend the polytropic dark halo model of \citet{saxton2008} to galaxy scales. We apply the model to fitting planetary nebula and stellar kinematics of elliptical galaxies. We present fits for a number of cases, including ``stars only'' models, ``stars + halo'' models with isotropic orbits and a constant global baryon fraction, and finally we also consider the issue of orbital anisotropy and a non-universal baryon fraction. \section{Method} \subsection{Spherical model and its solutions} We assume a fixed stellar mass distribution, with a projected surface density approximately described by the \citet{sersic1968} profile. S\'ersic-like profiles appear ubiquitous among observed galaxy spheroids, although the reasons why this form emerges in nature have not yet been demonstrated analytically. For the three-dimensional stellar density, we adopt the \citet{prugniel1997} spherical deprojection \citep[see also][]{limaneto1999,cotti1999,marquez2000}. \begin{equation} \rho_\bigstar(x) = \rho_{\rm e} x^{-p} \exp\left[{-b\left({x^{1/n}-1}\right)}\right] \ , \end{equation} with dimensionless coordinate $x\equiv r/r_{\rm e}$, and $(b,p)$ are constants depending on the index $n$. The scale radius (3D) is approximately equal to the observational, projected (2D) half-light radius, $r_{\rm e}\approx R_{\rm e}$. The stellar mass within some radius is given by \begin{equation} m_\bigstar(<x) = 4\pi n b^{n(p-3)} e^b \rho_{\rm e} r_{\rm e}^3 \ \Gamma[n(3-p),b x^{1/n}] \end{equation} involving lower incomplete gamma functions. Within this empirical stellar distribution, we introduce a dark matter halo, with a density distribution to be solved consistently with the global gravitational potential. The local equation of state is polytropic, \begin{equation} P_{\rm d} = s_{\rm d}\,\rho_{\rm d}^{\gamma_{\rm d}} \end{equation} with a pressure $P_{\rm d}$, density $\rho_{\rm d}$, pseudo-entropy $s_{\rm d}$ and adiabatic index $\gamma_{\rm d}\equiv 1 + 2/F_{\rm d}$. This is the natural behaviour expected for an isotropic medium comprising classical particles with $F_{\rm d}$ effective degrees of freedom. Such a condition is expected if dark matter is self-interacting \citep[e.g.][]{spergel2000,peebles2000,arbey2003,ahn2005,saxton2008}. An equivalent polytropic equation emerges if the dark halo is governed by Tsallis' non-extensive thermostatistics \citep[e.g.][]{tsallis1988,plastino1993,zavala2006,nunez2006}. If $-2<F_{\rm d}\le10$ then the dark halo may truncate at finite mass and radius ($R_{\rm h}$). For any particular model, we find this radius by assuming dark hydrostasis \begin{equation} {{dP_{\rm d}}\over{dr}} = -{{G\,m\,\rho_{\rm d}}\over{r^2}} \ , \end{equation} where we define the interior mass $m=m_\bigstar(r)+m_{\rm d}(r)$ for brevity. Computationally, it is more convenient to deal with the dark velocity dispersion (or temperature) $\sigma^2_{\rm d}=P_{\rm d}/\rho_{\rm d}$, which obeys \begin{equation} {{d\sigma^2_{\rm d}}\over{dr}} = -\left({2\over{F_{\rm d}+2}}\right) {{G\,m}\over{r^2}} \ . \end{equation} We take $s_{\rm d}$ as global constant, and numerically integrate the dark profile from the origin outwards. We assume non-singular conditions at the origin, \begin{equation} \left.{{d\rho_{\rm d}}\over{dr}}\right|_{r\rightarrow0}=0 \ , \end{equation} precluding the complication of a compact central mass and its sphere of influence. At some point the dark velocity dispersion becomes small (also $\rho_{\rm d}\rightarrow0$ if $F_{\rm d}>0$) and the integrator switches to use $\sigma^2_{\rm d}$ as the independent variable instead of $r$. Integration proceeds exactly to the point where $\sigma^2_{\rm d}=0$, i.e. the natural truncation radius $R_{\rm h}$. We call this the ``dark surface''. The low-mass region between the dark core (say where the log-slope of density is $-2$) and the dark surface shall be called the ``halo fringe.'' Once the total dark mass $M_{\rm d}$ and $R_{\rm h}$ are known, we can obtain the external gravitational potential, stellar pressure, potential energy and thermal energy profiles by integrating a Jeans equation (and associated ODEs for the mass profile and potential) inwards from infinity. If we define a stellar pressure in terms of the radial velocity dispersion of stars, $P_\bigstar\equiv\rho_\bigstar\sigma^2_\bigstar$ then we have the spherical Jeans equation \begin{equation} {{dP_\bigstar}\over{dr}} = -{{G\,m\,\rho_\bigstar}\over{r^2}} -{{2\beta P_\bigstar}\over{r}} \label{eq.jeans} \end{equation} where $\beta\equiv1-\sigma_\theta^2/\sigma^2_\bigstar$ is the local stellar velocity anisotropy. When integrating (\ref{eq.jeans}) far outside the dark halo, we change the independent variable to $u\equiv1/r$. Having obtained all the relevant constants of integration at the dark surface $R_{\rm h}$, we can easily evaluate profiles of the stellar and dark components in both the interior and exterior. \subsection{Model specification and projection} In any theory with $F_{\rm d}$ dark matter degrees of freedom fixed, the pseudo-entropy $s_{\rm d}$ and central dark temperature are formally free parameters. When discussing or fitting families of physically related models, it is more useful to prescribe the stellar mass fraction within some radius, $\mu_\bigstar(r) = m_\bigstar/(m_\bigstar+m_{\rm d})$, or the dark fraction $\mu_{\rm d}=1-\mu_\bigstar$. If galaxies retain all their primordial endowments of matter and darkness, then total stellar value would approximate the cosmic baryon fraction $\mu_\bigstar(\infty)\approx0.16$\footnote{We note that the latest WMAP-5 results give a best fit for a cosmological baryon fraction of $\sim$0.2 \citep{wmap5}. We refer the reader to \S\S3.4 for an estimate of a change in this number.} When seeking particular numerical solutions, we shall attain a chosen $\mu_\bigstar(\infty)$ by fixing the dark entropy $s_{\rm d}$ and adjusting the central $\sigma^2_{\rm d}$ iteratively. Then (for a given stellar background) there is an implicit relation between $s_{\rm d}$ and the dark truncation radius $R_{\rm h}$. Model observables are calculated by integrating suitable 3D stellar properties at radii $r=\sqrt{z^2+b^2}$ where $b$ is the projected radius on the sky plane and $z$ is the line-of-sight coordinate. The column density of stars is simply \begin{equation} \Sigma_\bigstar = 2\,\int_0^\infty \rho_\bigstar\,dz \ , \label{eq.column} \end{equation} and the sightline integrated pressure is \begin{equation} \Pi_\bigstar = 2\,\int_0^\infty P_\bigstar \left[{(1-\beta)+\beta{{z^2}\over{r^2}} }\right]\,dz \ . \label{eq.starpress} \end{equation} The mean line-of-sight velocity dispersion is then $\sigma_{\rm los}=\sqrt{\Pi_\bigstar/\Sigma_\bigstar}$. The orbital anisotropy is crucial to the stellar observables. We will mainly consider models with perfect isotropy ($\beta=0$ everywhere). In later sections (\S\ref{s.anisotropy}), we will consider anisotropy varying radially according to the Osipkov--Merritt model \citep{osipkov1979,merritt1985}, \begin{equation} \beta(r) = {{r^2}\over{r^2 + a^2}} \label{eq.OM} \end{equation} with the scale radius $a$ treated as a fitting parameter. In practice, integration of the stellar profile to radii at large orders of magnitude eventually samples a tenuous, hot region where $\rho_\bigstar$ drops much faster than $P_\bigstar$. Formally, the integrated velocity dispersion rises greatly. Physically, this contribution should not occur, because the densities imply fewer than one star within the relevant volume. Therefore we cut off the integrals (\ref{eq.column}) and (\ref{eq.starpress}) at some three-dimensional radius far outside the kinematic tracers of a particular galaxy. Calculations are performed in natural units of the S\'ersic profile: $\rho_{\rm e}=R_{\rm e}=G=1$. A numerical minimal-$\chi^2$ fitting procedure automatically yields the normalisation for a particular data set. This auto-normalisation counts as an extra free parameter, decrementing the ``degrees of freedom'' in the reduced $\chi^2$ scores. \section{Results} We fit the kinematics of stars and planetary nebulae of a sample comprising twelve galaxies from \citet{coccato2009} as well as NGC~1400 and NGC~1407. These data are separated into major and minor axis components. We fit the axes separately, to obtain an indication of the applicability of our model's spherical assumption. Additionally, we fit kinematic data for two of the galaxies without decomposition along axes: for NGC~821 we extract data from \citet{romanowsky2003}. For NGC~3379 we use figure~11 from \citet{douglas2007}. The data for NGC~1400 and 1407 are taken from \citet{proctor2009}. Notice in this case the data are extracted over a circular aperture, so only one axis can be considered. We assume that the stellar mass component follows the profile of \cite{prugniel1997}, with the S\'ersic parameters shown in the left three columns of Table~\ref{table.stars.optima}. For nine of the galaxies, these parameters come from photometric fits to S\'ersic-only models by \citet[][their table 3]{coccato2009}. For NGC~3379 we take $n=4.7$ from \citet{douglas2007}. For NGC~5128 we assume $n=4$ and obtain $R_{\rm e}$ from the B-band photometric fit of \citet{dufour1979}. For NGC~5846, we convert S\'ersic parameters from \citet{mahdavi2005b}. For all galaxies, we keep the distances and luminosities tabulated by \citet{coccato2009}. Initially we attempt fits without a polytrope (``stars only'') to test the necessity of dark matter, and to provide benchmarks for the fitting qualities of halo models. \subsection{Fits with stars only} Previous modeling of PN kinematics have suggested a poverty of dark matter in the halos of some isolated elliptical galaxies \citep{mendez2001,romanowsky2003} or that the halos are less concentrated than popular collisionless dark matter cosmogonies predict \citep[e.g.][]{douglas2007,napolitano2009,coccato2009}. To provide baselines for comparison, we first fit the Coccato et~al. data with a purely S\'ersic model (using the published, photometrically obtained parameters) and no dark matter. Except for NGC~4697, the goodness of fit scores -- given by the reduced $\chi_r^2\equiv\chi^2/\nu$ -- are well above one, and therefore bad fits (see columns 4 and 5 of Table~\ref{table.stars.optima} and the solid lines in figure~\ref{fig.nodark.fits}). The model profiles of the projected velocity dispersion are smoothly declining outside a half-light radius, whereas the observational profiles may be flat, or they may dip and rise with radius (Figure~\ref{fig.nodark.fits}). Some of these complexities and departures from ideal Sersic mass profiles may be responsible for the formally poor fits. \subsubsection{Improved stellar fits} Using the photometric S\'ersic parameters tabulated by \citet{coccato2009}, some galaxies give much worse values for the reduced $\chi^2$ than others. In order to improve these cases, we try fitting isotropic S\'ersic models without dark matter, and adapt the parameters ($n,R_{\rm e})$ for a better kinematic fit, ($n',R_{\rm e}')$. In this search, we omit the planetary nebulae, relying on absorption line kinematics only. Table~\ref{table.stars.optima} gives the alternative fits for axes 1 and 2 of each galaxy, before and after kinematic fitting. The left columns are based on \citet{coccato2009} photometric parameters; the right columns (primed quantities) are our kinematic fits. We give the quality of fit $(\chi^2_r)_{\{1,2\}}$ and the total (stellar) mass $m_{\bigstar_{\{1,2\}}}$, where the subindex refers to the axis considered in the analysis. These fits are shown in figure~\ref{fig.nodark.fits} as dotted lines. In some cases we use subindex 0 to refer to data that were obtained within circular annuli. We use an ``amoeba'' algorithm to find better fitting parameters, $(n^\prime,R_{\rm e}^\prime)$, with considerably improved $(\chi^2_r)^\prime$ on each axis. For NGC~3377, 1407 and 5128 the amoeba runs away forever along a degenerate $\chi^2$ valley, heading towards $R_{\rm e}\rightarrow\infty$. In the majority of attempts at $(n',R_{\rm e}')$ kinematic fits, we find $R_{\rm e}$ drifting to large values beyond the data, resulting in unphysically large total stellar masses. NGC~4564 and 3608 have smaller $R_{\rm e}'$ than the photometric values. NGC~1023 is the only galaxy where the photometric and kinematic fitted $R_{\rm e}$ values are moderate and comparable. Disagreements of photometric and kinematic S\'ersic parameters might indicate spatially varying stellar mass/light ratio, or the existence of dark matter. Alternatively, the disagreements may be influenced by degeneracies between the parameters, and the incomplete, finite radial range of the data. \subsection{Fitted halos with isotropic orbits} For the \citet{coccato2009} galaxies, data are available along two principal axes. We fit these independently, using a stellar mass profile with S\'ersic model parameters adopted from the published stellar photometry (i.e. $n$ and $R_{\rm e}$ from table~\ref{table.stars.optima}). The central properties are tuned so that the dark halo and stars are globally in a chosen ratio. We explore cases with fixed $F_{\rm d}$, and vary the halo radius (via $s_{\rm d}$) to achieve the best fit. Comparing the fits for the two axes gives some indication of the quality of the spherical approximation. (It is an imprecise comparison however, since the minor axis data tend to span a shorter radial range, and are less restrictive.) For six galaxies there are formally satisfactory fits, with $\chi^2_r\la1$ (i.e. in NGC~1407, 4374, 4564, 4697, 5128 and 5846). For the rest, the minimal $\chi^2_r$ never drops below a few (formally unfavourable). This may be partly due to ideosyncratic stellar kinematic substructures or noisiness. To address this possibility, we repeat the kinematic fitting process with the stellar $\sigma_{\rm los}(r)$ profile smoothed to a best-fit cubic function in $\ln(r)$. PN data are left untouched. Smoothing the stellar kinematics does not shift the minima significantly, but lowers the attainable $\chi^2_r$. Thus the minima are sharper and deeper, but the landscape is unchanged at medium and high $\chi^2$ levels. Results for the smoothed fitting are shown in Figures~\ref{fig.isotropic} and \ref{fig.isotropic2}. The colours from violet to red represent dark degrees of freedom $F_{\rm d} = \{1,2,3,...7,8,9,9.5,9.9\}$. The PN velocities were assumed to be isotropic. We assume a universal cosmic fraction of dark and baryonic mass (i.e. stars for our galaxies) according to the WMAP5 cosmology \citep{wmap5}. In figure~\ref{fig.isotropic2} we show four galaxies for which we have data within circular apertures \citep{romanowsky2003,douglas2007,proctor2009}. The usual $\chi^2$ curve of a galaxy has two minima: a shallow dip around $R_{\rm h}\sim0.1R_{\rm e}$ to 1$R_{\rm e}$, and a deeper dip at radii outside the observations. The physical interpretation of the inner dip implies a dark halo more compact than the stellar distribution. This would be surprising in the conventional cosmogonies in which radiative cooling made the star-forming gas more concentrated than the dark component. Numerically, at inner-dip configurations the ``halo'' is taking over the role of the stars influencing the inner $\sigma_{\rm los}$ profile. The outer minima seem more physically plausible. Sometimes the outer dip is effectively a lower limit on $R_{\rm h}$ (e.g. for NGC~4494) but in other cases the fitting prefers a finite range (e.g. less than 100$R_{\rm e}$ for NGC~821). If the data are clear and widespread enough, the radial variation of the slope of $\rho_{\rm d}$ in the observed range is sufficient to distinguish $F_{\rm d}$ and extrapolate the truncation radius. For $F_{\rm d}<9$ the $\chi^2(R_{\rm h})$ curves tend towards the ``stars only'' level as $R_{\rm h}\rightarrow\infty$. A large-$R_{\rm h}$ halo (of fixed $\mu_{\rm d}$) is spread thinly and has negligible effect on observable kinematics. The convergence is closer for lower $F_{\rm d}$, as these models resemble the incompressible limit ($F_{\rm d}=0$) which has uniform density within $r<R_{\rm h}$. The highly compressible $F_{\rm d}\ga9$ cases behave quite differently, as discussed in \S\ref{s.features}. Figure~\ref{fig.chi2} shows the two-dimensional 90, 95 and 99\% confidence levels for six galaxies in our sample, in the parameter space spanned by the dark matter degrees of freedom ($F_{\rm d}$) and the halo truncation radius ($R_{\rm h}$). The crosses mark the best fit, and the arrow represents the region over which the velocity dispersions are fit. The above mentioned degeneracy seen in the 1D plots of figure~\ref{fig.isotropic} is actually a continuous one between $F_{\rm d}$ and $R_{\rm h}$. For three exceptional galaxies (NGC~1023, 3608, 4564) the inner minimum is very significant, or even dominant, in terms of $\chi^2$. Inspection of the $\sigma_{\rm los}$ profiles shows that the photometrically-derived S\'ersic model fits them especially poorly. The inner part of the stellar profile is unusually centrally peaked. NGC~3377 also has a peculiarly shaped $\sigma_{\rm los}$ profile, which may be responsible for the slight preference for $F_{\rm d}=9.9$ on the first axis, and for the failure to improve on the stellar-only baseline fit to the second axis. The modelling of these four misfit galaxies implies very small stellar mass to light ratios (in the $B$-band), typically $\Upsilon^B_\bigstar\la0.1\Upsilon^B_\odot$ (see NGC~$1023_{(1,2)}$, NGC~$3377_{(1)}$ and NGC~$4564_{(1,2)}$ in Table~\ref{table.mol}). This is implausible for the aged stellar populations of early-type galaxies. We deprecate the discussion of these particular galaxies. Except for these odd cases, the ``stars + halo'' fits tend to reject $F_{\rm d}\gtrsim 9$. Where the data are clear and smooth enough, the isotropic modelling generally prefers the dark halo to have $F_{\rm d}=8$ or fewer effective thermal degrees of freedom. There is a mild preference for the top end of the range. In models with $F_{\rm d}\ge9$ the central dark density is insensitive to variations of the outer halo radius ($R_{\rm h}$) or the halo specific entropy ($s_{\rm d}$). This inflexibility in fitting central conditions may cause the poor fits for $F_{\rm d}\ge 9$. The best fits favour halo radii of at least tens of $R_{\rm e}$. We now comment on the favoured isotropic models for specific galaxies. \subsubsection{NGC~821} The kinematics of planetary nebulae around NGC~821 show a radially declining velocity dispersion \citep{romanowsky2003}. This has been taken as evidence disfavouring the presence of a dark halo (unless the PN orbital anisotropy has significant radial variations) or else the halo is less concentrated than CDM expectations. On both axes of the \citet{coccato2009} data, we find that polytropic halo models fit significantly better than stars alone. The improvement is best on the first axis. The best fits ($\chi^2_r\approx1.7$) are for $F_{\rm d}\approx8$ and an outer radius $60\la R_{\rm h}/R_{\rm e}\la90$. These minima improve significantly upon the ``stars only'' fit, in the case of axis 1 by up to $\Delta\chi^2_r\sim$25. The upper left panel of Figure~\ref{fig.isotropic2} presents the equivalent reduced-$\chi^2$ scores obtained by fitting polytropic dark halos to the \citet{romanowsky2003} data for NGC~821. These data were annular in distribution (not divided onto axes) and less detailed than \citet{coccato2009}. The fits also prefer $F_{\rm d}\approx 8$, but the minima are shallower. For $F_{\rm d}=8$ the fit improves gradually for larger $R_{\rm h}$, but for $F_{\rm d}\le7$ there is an optimum at several tens of $R_{\rm e}$. For the dark mass fraction within 1$R_{\rm e}$, the most likely value is $\approx0.2-0.4$, comparable to the \citet{coccato2009} value ($\approx0.4$). The best fit stellar mass-to-light ratio $\Upsilon^B_\bigstar\sim 5\Upsilon^B_\odot$ is typical of metal-rich populations of age 5-6~Gyr \citep[see e.g.][]{bc03}. \subsubsection{NGC~1023} Our analysis favours a dark halo with $F_{\rm d}\approx9.9$ (near maximum) and small radius $R_{\rm h}<10 R_{\rm e}$. The poor halo fits for this galaxy are perhaps unsurprising given the poor fits to the ``stars only'' basic model. The stellar $\sigma_{\rm los}$ profile is more centrally peaked than the (photometric) S\'ersic model. It was observed \citep{noordermeer2008,coccato2009} that the inner regions $r<100\arcsec$) are affected by strong rotation. NGC~1023 is a SB0 galaxy with a fast rotating bar which would imply a maximal (i.e. baryon dominated) disc \citep{vpd02}. We omit NGC~1023 in the analysis. \subsubsection{NGC~1344} Fits to data from axis 2 are bad mainly because of the flat plateau in velocity dispersion in the inner region. Axis~1 gives an acceptable fit for $F_{\rm d}\sim$6, with a significant improvement over the stars-only model. According to \citet{teodorescu2005}, there is a 3.8$\times10^{11}$M$_\odot$ dark matter halo out to 3.5$R_{\rm e}$. A rough estimate of the baryon fraction -- assuming 90\% of the stellar mass is contained within 3.5$R_{\rm e}$ \citep{grah05} -- gives a baryon fraction within this region of 0.29, in agreement with our estimates ($f_B\sim 0.3$). \subsubsection{NGC~1400} This lenticular galaxy is one of the two bright members of a nearby group \citep[along with NGC~1407,][]{go93}. We extract the stellar kinematics from \citet{proctor2009}. We take their assumed distance, but use the B magnitude and S\'ersic model from \citet{spolaor2008}. Assuming a cosmic baryon fraction, none of the polytropic halo models fits as well as ``stars only.'' Perhaps this is attributable to the discrepancies between $\sigma_{\rm los}$ profiles from different observations \citep[noted by][]{proctor2009}. We also notice that NGC~1400 has a significant amount of rotation \citep[$v/\sigma\sim 0.5$ at $R_{\rm e}$,][]{bert94}, which could possibly explain its +0.3~dex residual with respect to the Fundamental Plane \citep{ps96}. \subsubsection{NGC~1407} This E0 galaxy is the brightest member of a nearby group \citep[including NGC~1400,][]{go93}. The stellar data were extracted from \citet{proctor2009}, and we refer again to \citet{spolaor2008} for the light profile. The polytropic model with cosmic dark matter fraction gives a good fit, with $F_{\rm d}\approx7$ and truncation radius $\sim$15$R_{\rm e}$. The halo models with $F_{\rm d}\ga9$ have significantly poorer fits than for low and medium effective degrees of freedom. \subsubsection{NGC~3377} Stellar data for this flattened (E5) galaxy show a pronounced ``S-bend'' in the $\sigma_{\rm los}$ profile. All the fits are formally bad. On the first axis, a halo with high $F_{\rm d}$ brings a slight -- but perhaps insignificant -- improvement. On the second axis, a dark halo does not improve the fit at all, and the $F_{\rm d}=9.5$ and 9.9 curves are worse than the others. We suspect that complicated orbital structure (especially the undulating projected kinematics) renders straightforward spherical models inadequate and inappliable to NGC~3377. Details of this profile have been attributed to the influence of a central massive object and discy dynamics \citep{gebhardt2003,copin04}. \citet{coccato2009} note a twist in the PN kinematics. We omit NGC~3377 from consideration in our conclusions. \subsubsection{NGC~3379} This galaxy is another of the notable systems that \citet{romanowsky2003} observed to have a radially declining $\sigma_{\rm los}$ profile. For the \citet{coccato2009} data on two axes, the dark halo does not improve the fits over a S\'ersic-only model. Dark matter is allowed if the halo has a large radius and low central concentration. Cases with $F_{\rm d}\ge9$ are much worse fits than for lower $F_{\rm d}$. For comparison with the split-axial results, we also fit the projected velocity dispersion data representing circular rings, which are extracted from figure~11 of \citet{douglas2007} \citep[who incorporated stellar data from][]{statler1999}. A purely stellar model fits poorly: if the orbits are isotropic then $\chi^2_r\approx4.1$. If dark matter is present in the cosmic fraction, then it improves the fitting somewhat: the best $\chi^2_r\approx$2 occurs for $F_{\rm d}\approx7$ (see lower middle panel, Figure~\ref{fig.isotropic}). Isotropic models with $F_{\rm d}\ge9$ fit unacceptably worse than those with $F_{\rm d}\le8$. The simple S\'ersic model of the stellar mass may be inadequate to improve upon these fits. A bump in the stellar $\sigma_{\rm los}$ profile suggests the presence of kinematic substructure. \subsubsection{NGC~3608} Including a dark halo does not improve on stellar-only fits. The minima -- which have very high values of $\chi_r^2$ -- either imply a very compact halo ($R\sim$1$R_{\rm e}$) with $F_{\rm d}>9$ or an extended halo ($\sim 500$$R_{\rm e}$) with $F_{\rm d}=1$. We tentatively blame this outcome on the centrally peaky $\sigma_{\rm los}$ profile. If the S\'ersic profile describes the stellar mass distribution poorly, then the fitting routine compensates by seeking a high central dark density. \citet{coccato2009} remark that this galaxy had few PN detections, and possible contamination from NGC~3607 nearby. We omit NGC~3608 from further detailed examination. \subsubsection{NGC~4374} This E1 radio galaxy (Messier~84) is a member of the Virgo cluster. Fits to this galaxy favour the presence of a dark halo, with $F_{\rm d}\approx8$ and $R\approx$100 to 1000$R_{\rm e}$. Unusually, for the second axis, the $F_{\rm d}=9$ case improves on the basic ``stars only'' model (although $F_{\rm d}\approx8$ is still the best fit). \subsubsection{NGC~4494} This is another of the intriguing cases where \citet{romanowsky2003} found a declining velocity profile implying a dark matter halo with a very low density \citep{napolitano2009}. From data on both axes of \citet{coccato2009}, the fitting characteristics resemble those of NGC~3379: dark matter brings no improvement. A halo is allowed if its outer radius far exceeds the observed region. \subsubsection{NGC~4564} Models fitted to this E6 galaxy strongly prefer the presence of dark matter, but with radius of only a few $R_{\rm e}$. Perhaps this surprising conclusion is due to the centrally peaky stellar $\sigma_{\rm los}$ profile. The mass and orbital distributions may depart significantly from a S\'ersic sphere. \citet{coccato2009} indicate that there is a rotation curve resembling ``an S0 galaxy rather than ... an elliptical galaxy''. The unphysically low values of M/L and of the baryon fraction confirm the fit for this galaxy must invoke non-spherical models. We omit NGC~4564 from the analysis. \subsubsection{NGC~4697} This flattened elliptical (E6) galaxy has especially numerous PN measurements. Its velocity dispersion declines in the outskirts, and the degree of rotation appears to decline outside $R_{\rm e}$ \citep{mendez2001,mendez2008,sambhus2006,delorenzi2008a,coccato2009}. Despite the non-sphericity, our fits to both axes favour the presence of a dark halo with $F_{\rm d}\approx7$ or $8$ consistently. The extrapolated outer radius is large, $R_{\rm h}\gg100R_{\rm e}$. Thus the characteristics of this galaxy are consistent with the other apparently good fits. \subsubsection{NGC~5128} For the famous nearby galaxy NGC~5128 (Centaurus~A) the axis 1 data strongly favour the presence of a dark halo with cosmically mean composition. The best fit indicates $F_{\rm d}\approx$6 and $R\approx 25R_{\rm e}$. This is fewer dark thermal degrees of freedom than in our other plausible fits. The halo radius also seems more compact than usual. The prominent dust lane implies a recent merger, which might have left persistent departures from spherical hydrostasis in the dark sector as well. (Persistent dark pulsations or rotation could possibly change the apparent $F_{\rm d}$ obtainable from inapplicably static kinematics.) Unfortunately the results from axis 2 are inconclusive due to insufficient data. \subsubsection{NGC~5846} NGC~5846 is the central galaxy of a nearby group, and it appears round (morphological type E0-E1) and non-rotating \citep{coccato2009}. We should have expected decisive fits. The S\'ersic parameters are debatable: by fitting stellar and PN photometry, \citet{coccato2009} find $n=12\pm2$, $R_{\rm e}=2903\pm192\arcsec$. However \citet{mahdavi2005b} fit the light profile with $n=3.95\pm0.05$ and $R_{\rm e}=69.5\pm5.5\arcsec$. We adopt the latter parameter pair (high S\'{e}rsic indices, normally paired with very large half-light radii are usually indicative of a fitting degeneracy). Our fits to axis 1 marginally favour a dark halo with a small outer radius, $R_{\rm h}\approx$ 7.8$R_{\rm e}$, and $F_{\rm d}\approx 1$. On axis 2, the S\'ersic-only fit is similar to a stellar$+$dark matter model. \subsection{Features of the dark matter profile} \label{s.features} In a Lane-Emden sphere unaffected by other mass components, the quasi-entropy can take any positive real value, and there is a one-to-one relation between the entropy and the natural outer truncation radius of the halo. We find that this is untrue for a halo perturbed by a \citet{prugniel1997} stellar component. For $F_{\rm d}\ga6$ or so, we numerically find that there is a minimum $s_{\rm d}$ at which the halo fringe becomes infinitely diffuse, failing to truncate. For many models the cosmic baryon fraction is unattainable beyond a particular range of the entropy (even if the halo is finite). The poor fitting by $F_{\rm d}\ga9$ halos has a related cause. When the degrees of freedom are numerous, the core is smaller and denser compared to the outskirts. This dense core tends to outweigh the central stellar density, providing a poor fit to the inner data. For $F_{\rm d}\ga9$ the dark mass always outweighs the stars within 1$R_{\rm e}$ (see upper lines in figure~\ref{fig.dark.fraction}). Furthermore, when $F_{\rm d}\ge9$ we find that inflating the halo (extending the fringe radially) does not reduce the central dark density significantly. For lower $F_{\rm d}$, the core expands somewhat as the truncation radius increases, and this enables fits with plausibly low dark densities within the half-light radius. Figure~\ref{fig.fits.ngc0821} illustrates the halo density profile for the best-fitting models of axis 1 of NGC~821 for the choices of $F_{\rm d}$ as shown in figure~\ref{fig.isotropic} (with the same colour coding). In our fits, this galaxy possesses a significant dark mass fraction within the half-light radius, and the dark density at 1$R_{\rm e}$ is comparable to the stellar density there. The $F_{\rm d}\ge9$ curves are structurally distinct: the dark density follows that of the stars closely for several decades in radius. The truncation radius is large, but nevertheless the central dark density is high. The more favourable fits, with $F_{\rm d}\le8$, involve halos that truncate within 100$R_{\rm e}$, and there is a much larger inner region where $\rho_{\rm d}>\rho_\bigstar$ (e.g. $r\approx$0.7$R_{\rm e}$ for $F_{\rm d}=8$). For all fits, the logarithmic slope of the dark matter halo density is around $-1$ around and within the half-light radius. This should not be mistaken for the $\rho_{\rm d}\propto r^{-1}$ cusps of collisionless N-body simulations. The innately flat core of the polytropic halo has been gravitationally pinched by the stellar mass distribution. The {\em total} mass density index wavers around $-2$ out to several $R_{\rm e}$, which may be consistent with implications from some gravitational lensing studies \citep{rus03,koop06,if08}. If $F_{\rm d}<9$ and if our models have typical mass concentrations then the polytropic halo of a truly isolated elliptical galaxy would eventually be found to decline more sharply than $\propto r^{-3}$, somewhere beyond $\sim10R_{\rm e}$. Satellite galaxies and correlated cosmological macrostructure in the vicinity may confuse the practical detection of this truncation. Figure~\ref{fig.scatter_dark} shows the scatter of the baryon fraction with respect to stellar masses for the best-fit models that adequately improve over the stellar-only baseline cases. Solid dots represent those galaxies with a good fit ($\chi^2_r<3$). The figure shows that within 1$R_{\rm e}$ most galaxies appear baryon dominated except for massive ones, mainly NGC~1407. At 5$R_{\rm e}$ the baryon fraction decreases in most cases to $\sim$20\%. For comparison, the analysis of strong gravitational lenses from SLACS \citep{bol08} is shown as a grey shade in the left panel -- the SDSS-selected lenses mainly probe the inner ($R\lesssim R_{\rm e}$) regions. There is also consistency with the estimates of \citet{cappellari2006} using the SAURON sample -- they quote a dark matter fraction of $\sim 30\%$ in the inner $R_{\rm e}$; or with the recent work of \citet{tor09}, with similar values. The grey cross and error bars in figure~\ref{fig.scatter_dark} ({\sl right}) correspond to the lensing analysis of the galaxy B1104-181 (at a redshift z=0.73) from \citet{ferreras2005}, which was observed within $R<3.7R_{\rm e}$ (limited by the position of the lensed images of the background source, which controls the accuracy of the lensing estimates). \subsection{Stellar brightness and mass fraction} Though we have only fourteen galaxies (and not all of them with satisfactory fits) it is interesting to consider their apparent properties as a population. Figure~\ref{fig.scatter_star} shows the relation between $B$-band stellar mass-to-light ratio ($\Upsilon_\bigstar$) and the total stellar masses ($m_\bigstar$), which are obtained jointly from the fitting process. Those galaxies with a good fit ($\chi^2_r<3$) are shown as solid dots. In grey we also show the {\sl dynamical} M/L values from \citet{vdm07} for the galaxies in our sample. Those values are typically obtained within $R_{\rm e}$, and agree rather well with our estimates using a polytropic dark matter halo. For comparison with stellar populations, we also show in the right side of figure~\ref{fig.scatter_star} the expected stellar M/L for a couple of synthetic stellar populations from the CB07 models \citep[e.g.][]{bruz07} for a \citet{chab03} IMF. Two metallicities are considered, as labelled. We find that the more massive galaxies tend to have older stellar populations with ages compatible with more detailed work based on equivalent widths \citep[see e.g.][]{tm05,benyam} Among our standard calculations, NGC~821 shows perhaps the most convincing evidence for a dark halo of cosmic mass fraction. In a set of exploratory calculations, we vary the assumed ratio of stellar to dark mass for this particular galaxy. As illustrated in Figure~\ref{fig.n0821.fstar}, the stellar mass fraction $\mu_\bigstar(\infty)$ does have some effect. If the galaxy is richer in dark matter than the cosmic mean, then the best model shifts towards slightly higher $F_{\rm d}$, e.g. for $\mu_\bigstar=0.05$ -- a value motivated by the latest estimates of the baryon fraction in galaxy halos \citep[see e.g. ][]{moster09,guo09} -- the optimum is around $F_{\rm d}\approx9$ with $R_{\rm h}\approx457R_{\rm e}$. If the galaxy is poor in dark matter then $F_{\rm d}$ and $R_{\rm h}$ both shift to lower values. As $\mu_\bigstar\rightarrow1$, the $\chi^2$ curves collapse trivially to the poor fit of the ``stars only'' model. The best $\chi^2$ indicates moderate depletion of dark matter, e.g. when fixing $\mu_\bigstar=0.30$ we find $\chi^2_r=0.98$ and $R_{\rm h}\approx4.94R_{\rm e}$ for $F_{\rm d}=3$. This would imply halo truncation within the span of the observed planetary nebulae. Models with $\mu_\bigstar(\infty)\la0.2$ do not imply such a peculiarity. \subsection{Effect of radial orbital anisotropy} \label{s.anisotropy} To explore the role of orbital anisotropy, we try re-fitting some of the galaxies assuming the Osipkov-Merritt profile for $\beta(r)$ (see equation~\ref{eq.OM}). In this model, stellar orbits are isotropic in the centre, and increasingly radial in the fringe beyond a scale radius $a$ \citep{osipkov1979,merritt1985}. For ``stars only'' fits the OM radial anisotropy profile provides negligible improvement if the scale radius $a$ is large (approaching isotropic models), and for small $a$ it makes the $\chi^2$ scores significantly worse. Therefore radial anisotropy in the outskirts seems unlikely to cure the apparent dark deficit (or low halo concentration) of cases like NGC~4494. In this sense, ``stars alone'' models may be insufficient without non-spherical effects. Figure~\ref{fig.n0821.OM} shows the effect of Osipkov-Merritt radial orbital anisotropy profiles upon ``stars + halo'' models of NGC~821 (axis 1). NGC~821 is one of the cases with a clear need for a dark halo and this persists when the fringe is radially anisotropic. The $\chi^2$ curves still discriminate between $F_{\rm d}$ sharply. The curves prefer low $F_{\rm d}$ if $a<2$; for $a=2$ the optimum is $F_{\rm d}=3$; for $a=3$ it is $F_{\rm d}=6$. Thus the Osipkov-Merritt radial anisotropy model favours small $F_{\rm d}$ if the scale radius is comparable to the half-light radius, while isotropic models favour $F_{\rm d}\approx8$. The effects of the OM model are similar for other galaxies, such as NGC~3379. Overall $\chi^2$ levels are lower for larger $a$, and thus the isotropic limit is formally favourable. To convincingly constrain the amount and functional form of the anisotropy profile, it would be necessary to fit more detailed models with axisymmetry or fitting velocity kurtosis data. Such complexity is beyond the scope of our present paper. \subsection{Exotic models: $-2<F_{\rm d}<0$} It is worth mentioning the mathematical existence of a range of more exotic polytropes, with negative degrees of freedom. Models with $-2<F_{\rm d}<0$ and negative pressure ($P_{\rm d}<0$, $s_{\rm d}<0$) describe a ``generalised Chaplygin gas,'' which has been proposed to unify dark matter and dark energy \citep[e.g][]{bento2002,bilic2002,bertolami2004}. Polytropes with this composition have infinite radius and mass, so they need arbitrary truncation at some background density \citep[e.g.][]{bertolami2005}. Thus it is unclear how to naturally constrain the dark mass fraction in Chaplygin models of galaxies. However, if the pressure is positive, then the mass and radius of the halo are finite \citep{viala1974b,viala1974,kimura1981,lipscombe2008}. In this ``anti-Chaplygin gas'' halo, density is minimal at the origin, and increases to infinity at the outer radius $R_{\rm h}$ where temperature drops to zero. (Farther out, the dark density stays zero, as it does outside normal polytropes with $0<F_{\rm d}<10$.) These peculiar bubble-halo models are calculable, though the infinitely dense surface causes difficulty for the numerical integrators that project $\sigma_{\rm los}$. We performed a sparse set of exploratory calculations in this regime, and find a continuation of the trends in $\chi^2(R_{\rm h},F_{\rm d})$ already seen at the low end of the $F_{\rm d}>0$ domain. For NGC~821, the $\chi^2$ optimum rises (poorer fitting) and continues smoothly shifting to smaller $R_{\rm h}$ as $F_{\rm d}\searrow-2$. The limit of $F_{\rm d}=-2$ describes an isobaric condition. In this case, hydrostasis requires the gravitational field to vanish everywhere. There are no bound, finite, static, spherical, adiabatic solutions. The choice of $F=-2$ and $P_{\rm d}=0$ could describe a collisionless medium, but the dark density and entropy $s_{\rm d}$ would then be arbitrarily spatially variable (which is a complication beyond the scope of this paper). \section{Conclusions} We have fitted published kinematics of fourteen diverse elliptical galaxies, assuming a S\'ersic stellar mass profile plus a spherical polytropic dark matter halo that is distributed self-consistently in the shared gravitational field. This non-standard model is justified by our current lack of knowledge about the nature of the dark matter particle. A non-zero cross section for non-gravitational interactions or internal degrees of freedom will justify this interpretation. Furthermore, studies on the effect of gravitation on an ensemble of particles suggest a polytropic equation of state is required even if we are dealing with a non-interacting, elementary particle \citep{tsallis1988}. Our fits imply that more massive galaxies have higher stellar mass-to-light ratios, and higher central fractions of dark matter. We also impose a constraint on the degrees of freedom for the dark matter particle, $F_{\rm d}\gtrsim 6$, in agreement with a previous study focussed on galaxy clusters \citep{saxton2008}. The favoured $F_{\rm d}$ increases to higher values if elliptical galaxies are poorer in dark matter (and the favoured $F_{\rm d}$ is smaller if baryons are lost). Dark matter fractions well above or below the cosmic fractions may be explicable in several ways. Dark matter may be lost via ablation or tidal stripping, especially in dense environments. Baryonic matter might be lost due to blowout from stellar or AGN feedback, or from inefficient cooling (at the massive end), letting the baryons escape as warm or hot IGM/ICM unbound to the galaxy. The tentative preference for six to eight dark matter thermal degrees of freedom is interesting when compared to circumstantial evidence from galaxy clusters. In their analyses of galaxy clusters comprising polytropic dark matter and radiative gas, \citet{saxton2008} found that the condition for stationary solutions implies a lower limit on the central mass. Compatibility with the scale of observed supermassive black holes (as well as the dark matter core radii of clusters) implies a constraint of $7\la F_{\rm d}\la10$. (The upper limit is necessary to form any finite self-gravitating body.) Taken together, the evidence from elliptical galaxies and galaxy clusters therefore implies that dark matter has $F_{\rm d}\sim7$ to 8 (if it is the same substance in both contexts). Half of the available galaxies are well fit by the polytropic halo model, despite its serious physical simplifications. Some galaxies may require tailored, non-spherical modelling or special treatment of the kinematic relics of their ideosyncratic merger histories. We have assumed that the adiabatic dark halo is well mixed and is non-rotating: this may be untrue for a galaxy that has merged less than a few dynamical times ago. Our present modelling precludes the possible effects of a central star cluster or giant black hole, either of which could induce a parsec-scale spike of dark and stellar matter (within some gravitational radius of influence). This could raise the central velocity dispersion. Allowing this feature to emerge in future models might improve fits to the three galaxies with centrally spiky $\sigma_{\rm los}(r)$ profiles. \section*{Acknowledgments} We thank: K.~Wu for encouragement and criticism. M.~Cappellari for a conversation prompting this work. N.~Napolitano and L.~Coccato for their provision of online data. \bibliographystyle{mn2e}
\section*{Introduction} Let $\g$ be a simple Lie algebra and let $\g=\n^+\oplus\h\oplus \n^-$ be a Cartan decomposition. For a dominant integral $\la$ we denote by $V(\la)$ the irreducible $\g$-module with highest weight $\la$. Fix a highest weight vector $v_\la\in V(\la)$. Then $V(\la)=\U(\n^-)v_\la$, where $\U(\n^-)$ denotes the universal enveloping algebra of $\n^-$. The degree filtration $\U(\n^-)_s$ on $\U(\n^-)$ is defined by: $$ \U(\n^-)_s=\mathrm{span}\{x_1\dots x_l:\ x_i\in\n^-, l\le s\}. $$ In particular, $\U(\n^-)_0=\C$ and $gr \U(\n^-)\simeq S(\n^-)$, where $S(\n^-)$ denotes the symmetric algebra over $\n^-$. The filtration $\U(\n^-)_s$ induces a filtration $V(\la)_s$ on $V(\la)$: $$ V(\la)_s=\U(\n^-)_sv_\la.$$ We call this filtration the PBW filtration. In this paper we study the associated graded space $gr V(\la)$ for $\g$ of type ${\tt A}_n$. So from now on we fix $\g=\msl_{n+1}$. Note that $gr V(\la)=S(\n^-)v_\la$ is a cyclic $S(\n^-)$-module, so we can write $$ gr V(\la)\simeq S(\n^-)/I(\la), $$ for some ideal $I(\la)\subset S(\n^-)$. For example, for any positive root $\al$ the power $f_\al^{(\la,\al)+1}$ of a root vector $f_\al\in\n^-_{-\alpha}$ belongs to $I(\la)$ since $f_\al^{(\la,\al)+1}v_\la=0$ in $V(\la)$. To describe $I(\la)$ explicitly, consider the action of the opposite subalgebra $\n^+$ on $V(\la)$. It is easy to see that $\n^+ V(\la)_s\hk V(\la)_s$, so we obtain the structure of an $U(\n^+)$-module on $gr V(\la)$ as well. We show: \vskip 5pt\noindent {\bf Theorem~A.}\ $I(\la)=S(\n^-)\left(\U(\n^+)\circ \mathrm{span}\{f_\al^{(\la,\al)+1}, \al>0\}\right)$. \vskip 5pt Theorem $A$ should be understood as a commutative analogue of the well-known description of $V(\lambda)$ as the quotient $$ V(\la)\simeq \U(\n^-)/\langle f_\al^{(\la,\al)+1}, \al>0 \rangle $$ (see for example \cite{H}). Our second problem (closely related to the first one) is to construct a monomial basis of $gr V(\la)$. The elements $\prod_{\al>0} f_\al^{s_\al} v_\la$ with $s_\al\ge 0$ obviously span $gr V(\la)$ (recall that the order in $\prod_{\al>0} f_\al^{s_\al}$ is not important since $f_\al$ are considered as elements of $S(\n^-)$). For each $\la$ we construct a set $S(\la)$ of multi-exponents $\bs=\{s_\al\}_{\al>0}$ such that the elements $$ f^\bs v_\la=\prod_{\al>0} f_\al^{s_\al}v_\la, \ \bs\in S(\la) $$ form a basis of $gr V(\la)$. To give a precise definition of $S(\la)$ we need to introduce the notion of a {\it Dyck path}, which occurs already in Vinberg's conjecture: Let $\al_1,\dots,\al_n$ be the set of simple roots for $\msl_{n+1}$. Then all positive roots are of the form $\al_{p,q}=\al_p + \dots + \al_q$ for some $1\le p\le q\le n$. We call a sequence \[ \bp=(\beta(0), \beta(1),\dots, \beta(k)), \ k\ge 0, \] of positive roots a Dyck path (or simply a path) if it satisfies the following conditions: either $k=0$, and then $\bp =(\al_i)$ for some simple root $\alpha_i$, or $k\ge 1$, and then $\beta(0)=\al_i$, $\beta(k)=\al_j$ for some $1\le i< j\le n$ and the elements in between obey the following recursion rule: $$ \text{if } \beta(s)=\al_{p,q} \qquad \text{ then } \qquad \beta(s+1)=\al_{p,q+1} \text{ or } \beta(s+1)=\al_{p+1,q}. $$ Denote by $\D$ the set of all Dyck paths. For a dominant weight $\la=\sum_{i=1}^n m_i\omega_i$ let $P(\lambda)\subset \R^{\frac{1}{2}n(n+1)}_{\ge 0}$ be the polytope \begin{equation} \label{polytopeequation} P(\lambda):=\bigg\{(r_ \al)_{\al> 0}\mid \begin{array}{l} \forall\bp\in\D: \text{ If }\beta(0)=\al_i,\beta(k)=\al_j, \text{ then }\\ r_{\beta(0)} + \dots + r_{\beta(k)}\le m_i + \dots + m_j\\ \end{array}\bigg\}, \end{equation} and let $S(\la)$ be the set of integral points in $P(\lambda)$. We show: \vskip 5pt\noindent {\bf Theorem~B}.\ {\it The set of elements $f^\bs v_\la$, $\bs\in S(\la)$, forms a basis of $gr V(\la)$.} \vskip 5pt For $\bs \in S(\lambda)$ define the weight $$ \operatorname{wt}(\bs) := \sum_{1 \leq j \leq k \leq n} s_{j,k} \alpha_{j,k}. $$ As an important application we obtain \begin{cor*} \begin{itemize} \item[] \item[i)] For each $\bs\in S(\la)$ fix an arbitrary order of factors $f_\al$ in the product $\prod_{\al >0} f_\al^{s_\al}$. Let $f^\bs=\prod_{\al >0} f_\al^{s_\al}$ be the ordered product. Then the elements $f^\bs v_\la$, $\bs\in S(\la)$, form a basis of $V(\la)$. \item[ii)] $\dim V(\lambda)=\sharp S(\la)$. \item[iii)] $char V(\lambda) =\sum_{\bs\in S(\lambda)} e^{\la-\operatorname{wt}(\bs)}$. \end{itemize} \end{cor*} We note that the order in the corollary above is important since we are back to the action of the (in general) not commutative enveloping algebra. We thus obtain a family of bases for irreducible $\msl_{n+1}$-modules. Motivated by a different background, the existence of these bases (with the same indexing set) was conjectured by Vinberg (see \cite{V}). Using completely different arguments, he proved the conjecture for ${\mathfrak{sl}}_4$, for ${\mathfrak{sp}}_4$ and ${\tt G}_2$. Note also that the data labeling the basis vectors is similar to that for the Gelfand-Tsetlin patterns (see \cite{GT}). However, these bases are very different from the GT basis. \begin{exam}\rm For $\g=\msl_3({\mathbb C})$, there are only three Dyck paths, the two of length 1 corresponding to the simple roots, and the path which involves all positive roots. In the following we write the elements of $P(\lambda)$ in a triangular form, where we put $r_1=r_{\alpha_1}$ and $r_2=r_{\alpha_2}$ in the first row and $r_{12}=r_{\alpha_1+\alpha_2}$ in the second row. For $\lambda=m_1\om_1+m_2\om_2$ the associated polytope is $$ P(\lambda)=\bigg\{\Skew(0:r_1,r_2|.5:r_{12})\in\R_{\ge 0}^3\mid \begin{array}{l} 0\le r_1\le m_1,0\le r_2\le m_2,\\ r_1+r_2+r_{12}\le m_1+m_2 \end{array}\bigg\}, $$ For the set of integral points we get for example $$ S(\omega_1)=\bigg\{\Skew(0:0,0|.5: 0),\Skew(0:1,0|.5: 0),\Skew(0:0,0|.5: 1)\bigg\},\ S(\omega_2)=\bigg\{\Skew(0:0,0|.5: 0),\Skew(0:0,1|.5: 0),\Skew(0:0,0|.5: 1)\bigg\}. $$ and $$ S(2\omega_1+\omega_2)=\bigg\{\Skew(0:0,0|.5: 0), \Skew(0:1,0|.5: 0),\Skew(0:2,0|.5: 0), \Skew(0:0,0|.5: 1),\Skew(0:0,0|.5: 2), \Skew(0:0,0|.5: 3), \Skew(0:1,0|.5: 1), \Skew(0:1,0|.5: 2), $$ $$ \Skew(0:2,0|.5: 1), \Skew(0:0,1|.5: 0),\Skew(0:1,1|.5: 0),\Skew(0:2,1|.5: 0), \Skew(0:0,1|.5: 1),\Skew(0:0,1|.5: 2),\Skew(0:1,1|.5: 1)\bigg\}. $$ \end{exam} \bigskip We finish the introduction with several remarks. The PBW filtration for representations of affine Kac-Moody algebras was considered in \cite{FFJMT}, \cite{F1}, \cite{F2}. It was shown that it has important applications in the representation theory of current and affine algebras and in mathematical physics. There exist special representations $V(\la)$ such that the operators $f^\bs$ consist only of mutually commuting root vectors, even before passing to $gr V(\la)$. These modules can be described via the theory of abelian radicals and turned out to be important in the theory of vertex operator algebras (see \cite{GG}, \cite{FFL}, \cite{FL}). Let $V_w(\la)\hk V(\la)$ be a Demazure module for some element $w$ from the Weyl group. For special choices of $w$ there exists a basis of $V_w(\la)$ similar to the one given in Theorem $B$. We conjecture that this should be true for all $w\in W$ and we will discuss this elsewhere. Finally we note that $gr V(\la)$ carries an additional grading on each weight space $V(\la)^\mu$ of $V(\la)$: $$ gr V(\la)^\mu=\bigoplus_{s\ge 0} gr_s V(\la)^\mu=\bigoplus_{s\ge 0} V(\la)^\mu_s/V(\la)^\mu_{s-1}. $$ The graded character of the weight space is the polynomial $$ p_{\lambda,\mu}(q):=\sum_{s\ge 0}(\dim V(\la)^\mu_s/V(\la)^\mu_{s-1})q^s. $$ Define the degree $\deg(\bs) := \sum_{1 \leq j \leq k \leq n} s_{j,k}$ for $\bs\in S(\lambda)$, and let $S(\lambda)^\mu$ be the subset of elements such that $\mu=\lambda-\operatorname{wt}(\bs)$. Then \begin{cor*} $p_{\lambda,\mu}(q)=\sum_{\bs\in S(\lambda)^\mu} q^{\deg \bs}$. \end{cor*} We note that our filtration is different from the Brylinski-Kostant filtration (see \cite{Br}, \cite{K}). Our paper is organized as follows:\\ In Section 1 we introduce notations and state the problems. Sections 2 and 3 are devoted to the proof of Theorem $B$ (see Theorem~\ref{basis}). In Section 2 we prove the spanning property and in Section 3 the linear independence. In Section 4 we summarize our constructions and prove Theorem $A$ (see Theorem~\ref{idealtheorem}). \section{Definitions} Let $R^+$ be the set of positive roots of $\msl_{n+1}$. Let $\al_i$, $\omega_i$ $i=1,\dots,n$ be the simple roots and the fundamental weights. All roots of $\msl_{n+1}$ are of the form $\al_p + \al_{p+1} +\dots + \al_q$ for some $1\le p\le q\le n$. In what follows we denote such a root by $\al_{p,q}$, for example $\al_i=\al_{i,i}$. Let $\msl_{n+1}=\n^+\oplus\h\oplus \n^-$ be the Cartan decomposition. Consider the increasing degree filtration on the universal enveloping algebra of $\U(\n^-)$: \begin{equation}\label{df} \U(\n^-)_s=\mathrm{span}\{x_1\dots x_l:\ x_i\in\n^-, l \le s \}, \end{equation} for example, $\U(\n^-)_0=\C \cdot 1$. For a dominant integral weight $\la=m_1\omega_1 + \dots + m_n\omega_n$ let $V(\la)$ be the corresponding irreducible highest weight $\msl_{n+1}$-module with a highest weight vector $v_\la$. Since $V(\la)=\U(\n^-)v_\la$, the filtration \eqref{df} induces an increasing filtration $V(\la)_s$ on $V(\la)$: \[ V(\la)_s=\U(\n^-)_s v_\la. \] We call this filtration the PBW filtration and study the associated graded space $gr V(\la)$. In the following lemma we describe some operators acting on $gr V(\la)$. Let $S(\n^-)$ denotes the symmetric algebra of $\n^-$. \begin{lem} The action of $\U(\n^-)$ on $V(\la)$ induces the structure of a $S(\n^-)$-module on $gr V(\la)$ and \[ gr(V(\la))=S(\n^-)v_\la. \] The action of $\U(\n^+)$ on $V(\la)$ induces the structure of a $U(\n^+)$-module on $gr V(\la)$. \end{lem} \begin{proof} The first statement is obviously true by the definition of the filtrations $\U(\n^-)_s$ and $V(\la)_s$. The inclusions $\U(\n^+) V(\la)_s\hk V(\la)_s$ imply the second statement. \end{proof} Our aims are: \begin{itemize} \item to describe $gr V(\la)$ as an $S(\n^-)$-module, i.e. describe the ideal $I(\la)\hk S(\n^-)$ such that $gr V(\la)\simeq S(\n^-)/I(\la)$; \item to find a basis of $gr V(\la)$. \end{itemize} The description of the ideal will be given in the last section. To describe the basis we recall the definition of the Dyck paths: \begin{dfn}\rm A {\it Dyck path} (or simply {\it a path}) is a sequence \[ \bp=(\beta(0), \beta(1),\dots, \beta(k)), \ k\ge 0 \] of positive roots satisfying the following conditions: \begin{itemize} \item[{\it i)}] If $k=0$, then $\bp$ is of the form $\bp =(\al_i)$ for some simple root $\alpha_i$; \item[{\it ii)}] If $k\ge 1$, then \begin{itemize} \item[{\it a)}] the first and last elements are simple roots. More precisely, $\beta(0)=\al_i$ and $\beta(k)=\al_j$ for some $1\le i< j\le n$; \item[{\it b)}] the elements in between obey the following recursion rule: If $\beta(s)=\al_{p,q}$ then the next element in the sequence is of the form either $\beta(s+1)=\al_{p,q+1}$ or $\beta(s+1)=\al_{p+1,q}.$ \end{itemize} \end{itemize} \end{dfn} \begin{exam} Here is an example for a path for $\msl_6$: \[ \bp =(\al_2,\al_2+\al_3,\al_2+\al_3+\al_4,\al_3+\al_4,\al_4,\al_4+\al_5,\al_5). \] \end{exam} \vskip 5pt\noindent For a multi-exponent $\bs=\{s_ \beta\}_{\beta>0}$, $s_ \beta\in\Z_{\ge 0}$, let $f^\bs$ be the element \[ f^\bs=\prod_{\beta\in R^+} f_ \beta^{s_ \beta}\in S({\mathfrak n}^-). \] \begin{dfn} For an integral dominant $\msl_{n+1}$-weight $\la=\sum_{i=1}^n m_i\omega_i$ let $S(\la)$ be the set of all multi-exponents $\bs=(s_ \beta)_{\beta\in R^+}\in{\mathbb Z}_{\ge 0}^{R^+}$ such that for all Dyck paths $\bp =(\beta(0),\dots, \beta(k))$ \begin{equation}\label{upperbound} s_{\beta(0)}+s_{\beta(1)}+\dots + s_{\beta(k)}\le m_i+m_{i+1}+\dots + m_j, \end{equation} where $\beta(0)=\al_i$ and $\beta(k)=\al_j$. \end{dfn} In the next two sections we prove the following theorem. \begin{thm}\label{basis} The set $f^\bs v_\la$, $\bs\in S(\la)$, form a basis of $gr V(\la)$. \end{thm} \proof In Section 2 we show that the elements $f^\bs v_\la$, $\bs\in S(\la)$, span $gr V(\la)$, see Theorem~\ref{linearindipencetheorem}. In Section 3 we show that the number $\sharp S(\la)$ is smaller than or equal to $\dim V(\lambda)$ (see Theorem~\ref{numberofelements}), which finishes the proof of Theorem~\ref{basis}. \qed \section{The spanning property}\label{span} The space $gr V(\la)$ is endowed with the structure of a cyclic $S(\n^-)$-module, i.e. $gr V(\la)=S(\n^-)v_\la$ and hence $gr V(\la)=S(\n^-)/I(\la)$, where $I(\la)$ is some ideal in $S(\n^-)$. Our goal in this section is to prove that the elements $f^\bs v_\la$, $\bs\in S(\la)$, span $gr V(\la)$. Let $\la=m_1\omega_1 + \dots + m_n\omega_n$. The strategy is as follows: $f_\al^{(\la,\al)+1}v_\la=0$ in $V(\la)$ for all positive roots $\al$, so for $\alpha=\al_i + \dots + \al_j$, $i\le j$ we have the relation \[ f_{\al_i + \dots + \al_j}^{m_i+\dots + m_j+1} \in I(\la). \] In addition we have the operators $e_\al$ acting on $gr V(\la)$, and $I(\la)$ is stable with respect to $e_\al$. By applying the operators $e_\al$ to $f_{\al_i + \dots + \al_j}^{m_i+\dots + m_j+1}$, we obtain new relations. We prove that these relations are enough to rewrite any vector $f^\bt v_\lambda$ as a linear combination of $f^\bs v_\lambda$ with $\bs\in S(\la)$. We start with some notations. For $1\le i < j\le n$ set \[ \al_{i,j}=\al_i+\dots +\al_j,\ s_{i,j}=s_{\al_{i,j}},\ f_{i,j}=f_{\al_{i,j}}, \] and for convenience we set $\al_{i,i}=\al_i$, $s_{i,i}=s_{\al_{i}}$ and $f_{i,i}=f_{\al_{i}}$. By the degree $\deg \bs$ of a multi-exponent we mean the degree of the corresponding monomial in $S({\mathfrak n}^-)$, i.e. $\deg \bs=\sum s_{i,j}$. We are going to define a {\it monomial order} on $S({\mathfrak n}^-)$. To begin with, we define a total order on the set of generators $ f_{i,j}$, $1\le i\le j\le n$. We say that $(i,j)\succ (k,l)$ if $i>k$ or if $i=k$ and $j>l$. Correspondingly we say that $f_{i,j}\succ f_{k,l}$ if $(i,j)\succ (k,l)$, so \[ f_{n,n}\succ f_{n-1,n}\succ f_{n-1,n-1}\succ f_{n-2,n}\succ\ldots \succ f_{2,3}\succ f_{2,2}\succ f_{1,n}\succ\ldots\succ f_{1,1}. \] We use the associated {\it homogeneous lexicographic ordering} on the set of monomials in these generators of $S({\mathfrak n}^-)$. We use the ``same" total order on the set of multi-exponents, i.e. $\bs\succ \bt$ if and only if $f^\bs\succ f^\bt$. More explicitly: for two multi-exponents $\bs$ and $\bt$ we write $\bs\succ\bt$: \begin{itemize} \item if $\deg \bs>\deg \bt$, \item if $\deg \bs = \deg \bt$ and there exist $1\le i_0\le j_0\le n$ such that $s_{i_0j_0}>t_{i_0j_0}$ and for $i>i_0$ and ($i=i_0$ and $j>j_0$) we have $s_{i,j}=t_{i,j}$. \end{itemize} \begin{prop}\label{straightening} Let $\bp=(p(0),\dots,p(k))$ be a Dyck path with $p(0)=\al_i$ and $p(k)=\al_j$. Let $\bs$ be a multi-exponent supported on $\bp$, i.e. $s_\al=0$ for $\al\notin\bp$. Assume further that \[ \sum_{l=0}^k s_{p(l)}>m_i+\dots +m_j. \] Then there exist some constants $c_{\bt}$ labeled by multi-exponents $\bt$ such that \begin{equation}\label{straighteninglaw} f^\bs +\sum_{\bt<\bs} c_\bt f^\bt \in I(\lambda) \end{equation} ($\bt$ does not have to be supported on $\bp$). \end{prop} \begin{rem}\rm We refer to \eqref{straighteninglaw} as a {\it straightening law} because it implies $$ f^\bs = - \sum_{\bt<\bs} c_\bt f^\bt \text{\ in\ }S({\mathfrak n}^-)/I(\lambda)\simeq gr V(\lambda). $$ \end{rem} \begin{proof} We start with the case $p(0)=\al_1$ and $p(k)=\al_n$ (so, $k=2n-2$). This assumption is just for convenience. In the general case one has $\bp$ with $p(0)=\al_i$, $p(k)=\al_j$ and one would start with the relation $f_{i,j}^{m_i+\dots +m_j+1} \in I(\lambda)$ instead of the relation $f_{1,n}^{m_1+\dots +m_n+1} \in I(\lambda)$ below. So from now on we assume without loss of generality that $p(0)=\al_1$ and $p(k)=\al_n$. Let $S_+({\mathfrak h}\oplus {\mathfrak n}^+)\subset S({\mathfrak h}\oplus {\mathfrak n}^+)$ be the maximal homogeneous ideal of polynomials without constant term. The adjoint action of $U({\mathfrak n}^+)$ on ${\mathfrak g}$ induces an action of $U({\mathfrak n}^+)$ on $S({\mathfrak g})$ and hence on $$ S({\mathfrak n}^-)\simeq S({\mathfrak g})/S({\mathfrak n}^-)S_+({\mathfrak h}\oplus{\mathfrak n}^+). $$ In the following we use the differential operators $\pa_\al$ defined by $$ \pa_\al f_\beta= \begin{cases} f_{\beta-\al},\ \text{ if } \beta-\al\in\triangle^+,\\ 0,\ \text{ otherwise}. \end{cases} $$ The operators $\pa_\al$ satisfy the property \[ \pa_\al f_\beta = c_{\al,\beta} (\text{ad}\, e_\al)(f_\beta), \] where $c_{\al,\beta}$ are some non-zero constants. In the following we use very often the following consequence: if $f_{\beta_1}\dots f_{\beta_l}\in I(\la)$, then for any $\al_1,\dots,\al_s$ \[ \pa_{\al_1}\dots\pa_{\al_s} f_{\beta_1}\dots f_{\beta_l}\in I(\la). \] Since $f_{1,n}^{m_1+\dots +m_n+1} v_\lambda=0$ in $gr V(\la)$, it follows that \[ f_{1,n}^{s_{p(0)}+\dots + s_{p(k)}} \in I(\lambda). \] Write $\pa_{i,j}$ for $\pa_{\al_{i,j}}$. For instance, we have \begin{equation}\label{pa1} \pa_{1,i} f_{1,j}=f_{i+1,j},\ \pa_{j,n}f_{i,n}=f_{i,j-1} \ \text{ for } 1\le i<j\le n. \end{equation} For $i,j=1,\dots,n$ set \[ s_{\bullet, j}=\sum_{i=1}^j s_{i,j},\quad s_{i,\bullet}=\sum_{j=i}^n s_{i,j}. \] We consider first the vector \begin{equation}\label{pape1} \pa_{n,n}^{s_{\bullet, n-1}}\pa_{n-1,n}^{s_{\bullet, n-2}}\dots \pa_{2,n}^{s_{\bullet, 1}} f_{1,n}^{s_{p(0)}+\dots + s_{p(k)}} \in I(\lambda). \end{equation} Because of the formulas in \eqref{pa1} we get: \[ \pa_{2n}^{s_{\bullet, 1}} f_{1,n}^{s_{p(0)}+\dots + s_{p(k)}}=c_1 f_{1,n}^{s_{p(0)}+\dots + s_{p(k)-s_{\bullet, 1}}}f_{1,1}^{s_{\bullet, 1}} \] for some nonzero constant $c_1$, and \[ \pa_{3n}^{s_{\bullet, 2}}\pa_{2n}^{s_{\bullet, 1}} f_{1,n}^{s_{p(0)}+\dots + s_{p(k)}} =c_2 f_{1,n}^{s_{p(0)}+\dots + s_{p(k)}-s_{\bullet 1}-s_{\bullet 2}}f_{1,1}^{s_{\bullet 1}}f_{1,2}^{s_{\bullet 2}} \] for some nonzero constant $c_2$ etc. Summarizing, the vector \eqref{pape1} is proportional (with a nonzero constant) to \[ f_{1,1}^{s_{\bullet, 1}}f_{1,2}^{s_{\bullet ,2}}\dots f_{1,n}^{s_{\bullet, n}} \in I(\lambda). \] To prove the proposition, we apply more differential operators to the monomial $f_{1,1}^{s_{\bullet ,1}}f_{1,2}^{s_{\bullet, 2}}\dots f_{1,n}^{s_{\bullet, n}}$. Consider the following element in $I(\lambda)\subset S({\mathfrak n}^-)$: \begin{equation}\label{Aoperator} A=\pa_{1,1}^{s_{2,\bullet}}\pa_{1,2}^{s_{3,\bullet}}\dots \pa_{1,n-1}^{s_{n,\bullet}} f_{1,1}^{s_{\bullet ,1}}f_{1,2}^{s_{\bullet, 2}}\dots f_{1,n}^{s_{\bullet,n}}. \end{equation} {\bf We claim:} \begin{equation}\label{Aoperatorequation} A =\sum_{\bt\le \bs} c_\bt f^\bt \text{\ for some $c_\bs\ne 0$}. \end{equation} Now $A \in I(\lambda)$ by construction, so the claim proves the proposition. \vskip 5pt\noindent {\it Proof of the claim:} In order to prove the claim we need to introduce some more notation. For $j=1,\ldots,n-1$ set \begin{equation}\label{Ajoperator} A_j=\pa_{1,j}^{s_{j+1,\bullet}}\pa_{1,j+1}^{s_{j+2,\bullet}}\dots \pa_{1,n-1}^{s_{n,\bullet}} f_{1,1}^{s_{\bullet, 1}}f_{1,2}^{s_{\bullet, 2}}\dots f_{1,n}^{s_{\bullet, n}}, \end{equation} so $A_1=A$. To start an inductive procedure, we begin with $A_{n-1}$: $$ A_{n-1}=\pa_{1,n-1}^{s_{n,\bullet}} f_{1,1}^{s_{\bullet, 1}}f_{1,2}^{s_{\bullet, 2}}\dots f_{1,n}^{s_{\bullet, n}}. $$ Now $s_{n,\bullet}=s_{n,n}$ and $\pa_{1,n-1}f_{1,j}=0$ except for $j=n$, so \begin{equation}\label{Aoperatorequation1} A_{n-1}= c f_{1,1}^{s_{\bullet, 1}}f_{1,2}^{s_{\bullet, 2}}\dots f_{1,n}^{s_{\bullet, n}-s_{n,n}}f_{n,n}^{s_{n,n}}, \end{equation} for some nonzero constant $c$. The proof will now proceed by decreasing induction. Since the induction procedure is quite involved and the initial step does not reflect the problems occurring in the procedure, we discuss for convenience the case $A_{n-2}$ separately. Consider $A_{n-2}$, up to a nonzero constant we have: $$ A_{n-2}=\pa_{1,n-2}^{s_{n-1,\bullet}} f_{1,1}^{s_{\bullet, 1}}f_{1,2}^{s_{\bullet, 2}}\dots f_{1,n}^{s_{\bullet, n}-s_{n,n}}f_{n,n}^{s_{n,n}}. $$ Now $\pa_{1,n-2}f_{1,j}=0$ except for $j=n-1,n$, and $\pa_{1,n-2}f_{n,n}=0$, so $$ A_{n-2}=\sum_{\ell=0}^{s_{n-1 ,\bullet}} c_\ell f_{1,1}^{s_{\bullet, 1}}f_{1,2}^{s_{\bullet, 2}}\dots f_{1,n-1}^{s_{\bullet, n-1}-s_{n-1,\bullet}+\ell}f_{1,n}^{s_{\bullet, n}-s_{n,n}-\ell} f_{n-1,n-1}^{s_{n-1,\bullet}-\ell}f_{n-1,n}^{\ell}f_{n,n}^{s_{n,n}}. $$ We need to control which powers $f_{n-1,n}^{\ell}$ can occur. Recall that $\bs$ has support in $\bp$. If $\al_{n-1}\not\in \bp$, then $s_{n-1,n-1}=0$ and $s_{n-1,\bullet}=s_{n-1,n}$, so $f_{n-1,n}^{s_{n-1,n}}$ is the highest power occurring in the sum. Next suppose $\al_{n-1}\in \bp$. This implies $\al_{j,n}\not \in \bp$ unless $j=n-1$ or $n$. Since $\bs$ has support in $\bp$, this implies $$ s_{\bullet, n}=s_{1,n}+\ldots + s_{n-1,n}+s_{n,n}=s_{n-1,n}+s_{n,n}, $$ and hence again the highest power of $f_{n-1,n}$ which can occur is $f_{n-1,n}^{s_{n-1,n}}$, and the coefficient is nonzero. So we can write \begin{equation}\label{Aoperatorequation2} A_{n-2}=\sum_{\ell=0}^{s_{n-1,n}} c_\ell f_{1,1}^{s_{\bullet, 1} \dots f_{1,n-1}^{s_{\bullet, n-1}-s_{n-1,\bullet}+\ell}f_{1,n}^{s_{\bullet, n}-s_{n,n}-\ell} f_{n-1,n-1}^{s_{n-1,\bullet}-\ell}f_{n-1,n}^{\ell}f_{n,n}^{s_{n,n}}. \end{equation} For the inductive procedure we make the following assumption: $A_{j}$ is a sum of monomials of the form \begin{equation}\label{Aoperatorequation3} \underbrace{f_{1,1}^{s_{\bullet, 1}}\ldots f_{1,j}^{s_{\bullet, j}}f_{1,j+1}^{s_{\bullet, j+1}-*}\ldots f_{1,n}^{s_{\bullet, n}-*}}_X \underbrace{f_{j+1,j+1}^{t_{j+1,j+1}}f_{j+1,j+2}^{t_{j+1,j+2}}\dots f_{n-1,n}^{t_{n-1,n}}f_{n,n}^{t_{n,n}}}_Y \end{equation} having the following properties: \begin{itemize} \item[{\it i)}] With respect to the homogeneous lexicographic ordering, all the multi-exponents of the summands, except one, are strictly smaller than $\bs$. \item[{\it ii)}] More precisely, there exists a pair $(k_0,\ell_0)$ such that $k_0\ge j+1$, $s_{k_0 \ell_0}>t_{k_0 \ell_0}$ and $s_{k\ell}=t_{k\ell}$ for all $k>k_0$ and all pairs $(k_0.\ell)$ such that $\ell>\ell_0$. \item[{\it iii)}] The only exception is the summand such that $t_{\ell,m}=s_{\ell,m}$ for all $\ell\ge j+1$ and all $m$. \end{itemize} The calculations above show that this assumption holds for $A_{n-1}$ and $A_{n-2}$. We come now to the induction procedure and we consider $A_{j-1}= \pa_{1,j-1}^{s_{j,\bullet}}A_{j}$. Note that $\pa_{1,j-1} f_{1,\ell}=0$ except for $\ell \ge j$, and in this case we have $\pa_{1,j-1} f_{1,\ell}=f_{j,\ell}$. Furthermore, $\pa_{1,j-1} f_{k,\ell}=0$ for $k\ge j+1$, so applying $\pa_{1,j-1}$ to a summand of the form in (\ref{Aoperatorequation3}) does not change the $Y$-part in (\ref{Aoperatorequation3}). Summarizing, applying $\pa_{1,j-1}^{s_{j,\bullet}}$ to a summand of the form in (\ref{Aoperatorequation3}) gives a sum of monomials of the form \begin{equation}\label{Aoperatorequation4} \underbrace{f_{1,1}^{s_{\bullet ,1}}\ldots f_{1,j-1}^{s_{\bullet, j-1}}f_{1,j}^{s_{\bullet,j}-*}\ldots f_{1,n}^{s_{\bullet,n}-*}}_{X'} \underbrace{f_{j,j}^{t_{j,j}}\ldots f_{j,n}^{t_{j,n}} }_Z \underbrace{f_{j+1,j+1}^{t_{j+1,j+1}}f_{j+1,j+2}^{t_{j+1,j+2}}\dot f_{n,n}^{t_{n,n}}}_Y. \end{equation} We have to show that these summands satisfy again the conditions ${\it i)}$--${\it iii)}$ above (but now for the index $(j-1)$). If we start in (\ref{Aoperatorequation3}) with a summand which is not the maximal summand, but such that {\it i)} and {\it ii)} hold for the index $j$, then the same holds obviously also for the index $(j-1)$ for all summands in (\ref{Aoperatorequation4}) because the $Y$-part remains unchanged. So it remains to investigate the summands of the form (\ref{Aoperatorequation4}) obtained by applying $\pa_{1j-1}^{s_{j,\bullet}}$ to the only summand in (\ref{Aoperatorequation3}) satisfying {\it iii)}. To formalize the arguments used in the calculation for $A_{n-2}$ we need the following notation. Let $1\le k_1\le k_2\le\dots\le k_n\le n$ be numbers defined by \[ k_i=\max\{j:\ \al_{i,j}\in\bp\}. \] For convenience we set $k_0=1$. \begin{exam} For $\bp=(\al_{11},\al_{12},\dots,\al_{1n},\al_{2n},\dots,\al_{n,n})$ we have $k_i=n$ for all $i=1,\ldots,n$. \end{exam} Since $\bs$ is supported on $\bp$ we have \begin{equation}\label{sumpath} s_{i,\bullet}=\sum_{\ell=k_{i-1}}^{k_i} s_{i,\ell},\ s_{\bullet, \ell}=\sum_{i:\ k_{i-1}\le \ell\le k_i} s_{i,\ell}. \end{equation} Suppose now that we have a summand of the form in (\ref{Aoperatorequation4}) obtained by applying $\pa_{1j-1}^{s_{j,\bullet}}$ to the only summand in (\ref{Aoperatorequation3}) satisfying {\it iii)}. Since the $Y$-part remains unchanged, this implies already $t_{n,n}=s_{n,n},\ldots, t_{j+1,j+1}=s_{j+1,j+1}$. Assume that we have already shown $t_{j,n}=s_{j,n},\ldots, t_{j,\ell_0+1}=s_{j,\ell_0+1}$, then we have to show that $t_{j,\ell_0}\le s_{j,\ell_0}$. We consider five cases: \begin{itemize} \item $\ell_0 > k_{j}$. In this case the root $\al_{j,\ell_0}$ is not in the support of $\bp$ and hence $s_{j,\ell_0}=0$. Since $\ell_0>k_{j}\ge k_{j-1}\ge\ldots\ge k_1$, for the same reason we have $s_{i,\ell_0}=0$ for $i\le j$. Recall that the power of $f_{1,\ell_0}$ in $A_{j-1}$ in (\ref{Ajoperator}) is equal to $s_{\bullet, \ell_0}$. Now $s_{\bullet, \ell_0}=\sum_{i>j} s_{i,\ell_0}$ by the discussion above, and hence $f_{1,\ell_0}^{s_{\bullet, \ell_0}}$ has already been transformed completely by the operators $\pa_{1,i}$, $i>j$, and hence $t_{j,\ell_0}=0=s_{j,\ell_0}$. \item $k_{j-1}<\ell_0\le k_{j}$. Since $\ell_0 > k_{j-1}\ge\ldots\ge k_1$, for the same reason as above we have $s_{i,\ell_0}=0$ for $i<j$, so $s_{\bullet, \ell_0}=\sum_{i\ge j} s_{i,\ell_0}$. The same arguments as above show that for the operator $\pa_{1,j-1}$ only the power $f_{1,\ell_0}^{s_{j,\ell_0}}$ is left to be transformed into a power of $f_{j,\ell_0}$, so necessarily $t_{j,\ell_0}\le s_{j,\ell_0}$. \item $k_{j-1}=\ell_0=k_{j}$. In this case $s_{j,\bullet}=s_{j,\ell_0}$ and thus the operator $\pa_{1,j-1}^{s_{j,\bullet}}=\pa_{1,j-1}^{s_{j,\ell_0}}$ can transform a power $f^*_{1,\ell_0}$ in $A_j$ only into a power $f^q_{j,\ell_0}$ with $q$ at most $s_{j,\ell_0}$. \item $k_{j-1}= \ell_0 < k_{j}$. In this case $s_{j,\bullet}=s_{j ,\ell_0}+s_{j ,\ell_0+1}+\ldots+s_{j ,k_j}$. Applying $\pa_{1,j-1}^{s_{j,\bullet}}$ to the only summand in (\ref{Aoperatorequation3}) satisfying {\it iii)}, the assumption $t_{j,n}=s_{j,n},\ldots, t_{j,\ell_0+1}=s_{j,\ell_0+1}$ implies that one has to apply $\pa_{1,j-1}^{s_{j,k_j}}$ to $f_{1,k_j}^*$ and $\pa_{1,j-1}^{s_{j,k_j-1}}$ to $f_{1,k_j-1}^*$ etc. to get the demanded powers of the root vectors. So for $f^*_{1,\ell_0}$ only the operator $\pa_{1,j-1}^{s_{j,\ell_0}}$ is left for transformations into a power of $f_{j,\ell_0}$ and hence $t_{j,\ell_0}\le s_{j,\ell_0}$. \item $\ell_0< k_{j-1}$. In this case $s_{j,\ell_0}=0$ because the root is not in the support. Since $t_{j,\ell }=s_{j,\ell}$ for $\ell>\ell_0$ and $s_{j,\ell}=0$ for $\ell \le \ell_0$ (same reason as above) we obtain \[ \pa_{1,j-1}^{s_{j,\bullet}}=\pa_{1,j-1}^{\sum_{\ell>\ell_0} s_{j,\ell}}. \] But by assumption we know that $\pa_{1,j-1}^{s_{j,\ell}}$ is needed to transform the power $f_{1,\ell}^{s_{j,\ell}}$ into $f_{j,\ell}^{s_{j,\ell}}$ for all $\ell >\ell_0$, so no power of $\pa_{1,j-1}$ is left and thus $t_{j,\ell_0}=0=s_{j,\ell_0}$. \end{itemize} It follows that all summands except one satisfy the conditions {\it i),ii)} above. The only exception is the term where the powers of the operator $\pa_{1,j-1}^{s_{j,\bullet}}$ are distributed as follows: $$ f_{1,1}^{s_{\bullet, 1}}... f_{1,j-1}^{s_{\bullet, j-1}} (\pa_{1,j-1}^{s_{j,j}}f_{1,j}^{s_{\bullet, j}}) (\pa_{1,j-1}^{s_{j,j+1}}f_{1,j+1}^{s_{\bullet, j+1}-*}) ... (\pa_{1,j-1}^{s_{j,n}} f_{1,n}^{s_{\bullet, n}-*}) f_{j+1,j+1}^{s_{j+1,j+1}}... f_{n,n}^{s_{n,n}}. $$ By construction, this term is nonzero and satisfies the condition {\it iii)}, which finishes the proof of the proposition. \end{proof} \begin{thm}\label{linearindipencetheorem} The elements $f^{\bs} v_\lambda$ with $\bs\in S(\la)$ span the module $gr V(\la)$. \end{thm} \begin{proof} The elements $f^\bs v_\la$, $\bs$ arbitrary multi-exponent, span $S({\mathfrak n}^-)/I(\lambda)\simeq gr V(\la)$. We use now the equation \eqref{straighteninglaw} in Proposition~\ref{straightening} as a straightening algorithm to express $f^\bt v_\la $, $\bt$ arbitrary, as a linear combination of elements $f^\bs v_\la$ such that $\bs\in S(\lambda)$. Let $\lambda=\sum_{i=1}^n m_i\omega_i$ and suppose $\bs\notin S(\la)$, then there exists a Dyck path $\bp=(p(0),\dots,p(k))$ with $p(0)=\al_i$, $p(k)=\al_j$ such that \[ \sum_{l=0}^k s_{p(l)} > m_i +\dots + m_j. \] We define a new multi-exponent $\bs'$ by setting \[\bs'_\al= \begin{cases} s_\al, \ \al\in\bp,\\ 0,\ otherwise. \end{cases} \] For the new multi-exponent $\bs'$ we still have $$ \sum_{l=0}^k s'_{p(l)} > m_i +\dots + m_j. $$ We can now apply Proposition~\ref{straightening} to $\bs'$ and conclude $$ f^{\bs'} =\sum_{\bs'>\bt'} c_{\bt'} f^{\bt'}\quad\text{in}\quad S({\mathfrak n}^-)/I(\lambda). $$ We get $f^\bs$ back as $f^\bs=f^{\bs'}\prod_{\beta\notin\bp} f_\beta^{s_\beta}$. For a multi-exponent $\bt'$ occurring in the sum with $c_{\bt'}\not=0$ set $f^\bt= f^{\bt'}\prod_{\beta\notin\bp} f_\beta^{s_\beta}$ and $c_\bt=c_{\bt'}$. Since we have a monomial order it follows: \begin{equation}\label{tprime3} f^\bs =f^{\bs'}\prod_{\beta\notin\bp} f_\beta^{s_\beta}=\sum_{\bs>\bt} c_\bt f^\bt \quad\text{in}\quad S({\mathfrak n}^-)/I(\lambda). \end{equation} The equation \eqref{tprime3} provides an algorithm to express $f^\bs$ in $S({\mathfrak n}^-)/I(\lambda)$ as a sum of elements of the desired form: if some of the $\bt$ are not elements of $S(\lambda)$, then we can repeat the procedure and express the $f^\bt$ in $S({\mathfrak n}^-)/I(\lambda)$ as a sum of $f^\br$ with $\br<\bt$. For the chosen ordering any strictly decreasing sequence of multi-exponents is finite, so after a finite number of steps one obtains an expression of the form $f^\bs=\sum c_\br f^\br$ in $S({\mathfrak n}^-)/I(\lambda)$ such that $\br\in S(\lambda)$ for all $\br$. \end{proof} \section{The linear independence}\label{li} In the following let $R_i$ denote the subset $$ R_i = \{ \alpha \in R^+ \; | \; (\omega_i, \alpha) = 1 \}. $$ We define for a dominant weight $\lambda \in P^+$ $$ R_{\lambda} = \{ \alpha \in R^+ \; | \; (\lambda, \alpha) > 0\}. $$ Recall that we use $\al_{i,j}$ as an abbreviation for $\al_i+\al_{i+1}+\ldots+\al_j$ (see Section \ref{span}). The set $R_i$ can then be described as $$ R_i = \{ \alpha_{j,k} \; | \; 1 \leq j \leq i \leq k \leq n \}. $$ We say a path $\bp$ has {\it color} $i$ if $\exists \; j$ s.t. $\beta(j) \in R_i$. Note that a path can have several different {colors}. To simplify the notation we often just write $(j,k)$ for the root $\alpha_{j,k}$ (if no confusion is possible). Let $\lambda = \sum_{j=1}^{n} m_j \omega_j$ and let $i$ be minimal such that $m_i \neq 0$. For $\bs \in S(\lambda)$, we denote $$R^{\bs}_i = \{ (j,k) \in R_i \; | \; s_{j,k} \neq 0 \}.$$ We define two different orders on $R$, a partial order ``$\leq$": $$ (j_1, k_1) \leq (j_2, k_2) \Leftrightarrow (j_1 \leq j_2 \; \wedge k_1 \leq k_2), $$ and a total order ``$\ll$'': $$(j_1, k_1) \ll (j_2, k_2) \Leftrightarrow \text{ if }(k_1 < k_2) \text{ or } (k_1 = k_2 \wedge j_1 < j_2).$$ By definition, ``$\ll$'' covers ``$\leq$''.\\ \begin{exam}\rm For $\g=\msl_4$ and $i =2$, the minimal element of $R_i$ with respect to both orders is $(1,2) = \alpha_{1,2} = \alpha_1 + \alpha_2$. Note that $ \alpha_1 + \alpha_2 + \alpha_3 \ll \alpha_2$, but the two are not comparable with respect to ``$\leq$". \end{exam} A tuple $\bs \in S(\lambda)$ will be considered as an ordered tuple with respect to the order ``$\ll$": $$ \bs = (s_{1,1}, s_{1,2}, s_{2,2}, s_{1,3}, s_{2,3}, s_{3,3}, \ldots , s_{n, n}). $$ The induced lexicographic order on $S(\lambda)$ is a total order which we again denote by ``$\ll$". \begin{rem}\rm The total order $\ll$ is different from the order $\prec$ used in Section~\ref{span}. \end{rem} \begin{exam}\rm For $\g=\msl_4$ let $\bs$ be defined by $$ s_{13} = 1, s_{22} = 1 \text{ and } s_{j,k} = 0 \text{ otherwise,} $$ and let $\bt$ be defined by $$ t_{12} = 1, t_{23} = 1 \text{ and } t_{j,k} = 0 \text{ otherwise. } $$ Then $\bs= (0,0,1,1,0,0)$ and $\bt=(0,1,0,0,1,0)$, and so $\bs \ll \bt$. \end{exam} \begin{definition} For $\bs \in S(\lambda)$ denote by $M^{\bs}_i$ the {\it set of minimal elements} in $R^{\bs}_i$ with respect to $\leq$. We denote by $\bm^{\bs}_i$ the tuple $m_{j,k} = 1$ if $(j,k) \in M^{\bs}_i$ and $m_{j,k} = 0$ otherwise. \end{definition} \begin{exam}\rm 1) If $R^{\bs}_i=R_i$, then $M^{\bs}_i=\{\al_{1,i}\}$. \par 2) If $R^{\bs}_i=\{\al_{i,i},\al_{i-1,i+1},\ldots,\al_{i-\ell,i+\ell}\}$ for some $\ell\le i$, then $M^{\bs}_i=R^{\bs}_i$. \end{exam} \begin{rem}\label{minimalelements} $1)$.\ For any multi-exponent $\bs$ we have $$ M^{\bs}_i=\{ \al_{j_l, k_l} \; | \; l=1, \ldots, m \} $$ for some $m$, and the indices have the property $$ 1\le j_1 < j_2 < \ldots < j_m\le i \le k_m< \dots < k_2 < k_1 \le n. $$ If $\bs\in S(\omega_i)$, then for the associated tuple $\bm_i^\bs$ we get: $\bm_i^\bs=\bs$.\\ \noindent $2)$.\ The sets $M^{\bs}_i$ satisfy the following important property: any Dyck path contains at most one element of $M^{\bs}_i$, because the elements of a Dyck path are linearly ordered with respect to ``$\ge$''. \end{rem} \begin{prop}\label{minimalset} For $\bs \in S(\lambda)$ let $M^{\bs}_i$ be the minimal set. Then $\bm^{\bs}_i \in S(\omega_i)$, and if $\bs'$ is such that $\bs = \bs' + \bm^{\bs}_i$, then $\bs' \in S(\lambda - \omega_i)$. \end{prop} \proof Note that $\bm^{\bs}_i \in S(\omega_i)$ by Remark~\ref{minimalelements}. Let $\bs'$ be such that $\bs = \bs' + \bm^{\bs}_i$. We claim that $\bs' \in S(\lambda - \omega_i)$. Let $\lambda=\sum_{j=i}^n m_j\om_j$. For a Dyck path $\bp$ let $q_\bp^\lambda=\sum_{j\, \text{color of}\, \bp} m_j$ be the upper bound for the defining inequality \eqref{upperbound} of $S(\lambda)$ associated to $\bp$. If $\bp$ is a Dyck path such that $i$ is not a color, then $q_\bp^\lambda= q_\bp^{\lambda-\om_i}$ and $s_{\beta}=s_{\beta}'$ for $\beta\not\in R_i$, so $\bs'$ satisfies the defining inequality for $S(\lambda - \omega_i)$ given by $\bp$. Let $\bp$ be a Dyck path of color $i$, so $q_\bp^{\lambda-\om_i}=q_\bp^\lambda-1$. If $\bp\cap M^{\bs}_i\not=\emptyset$, then $\sum_{(j,k) \in \bp} s_{j,k}' = \sum_{(j,k) \in \bp} s_{j,k}-1\le q_\bp^\lambda-1=q_\bp^{\lambda-\om_i}$, so $\bs'$ satisfies the defining inequality for $S(\lambda - \omega_i)$ given by $\bp$. Suppose now that $\bp$ is a Dyck path of color $i$ but $\bp\cap M^{\bs}_i=\emptyset$. Recall that the elements in $\operatorname{supp} \bp$ are linearly ordered. Let $\al_{l,m}$ be the minimal element in $R^{\bs}_i \cap \operatorname{supp} \bp$. Since $i$ is minimal such that $m_i>0$, note that $s_\beta=0$ for all $\beta\in\operatorname{supp}\bp$ be such that $\beta<\al_{l,m}$. By assumption, $\al_{l,m}\not\in M^{\bs}_i$, so let $\al_{r,t}\in M^{\bs}_i$ such that $\al_{r,t}<\al_{l,m}$. Let $\tilde\bp$ be the Dyck path $$ (\al_{r,r},\al_{r,r+1},\ldots,\al_{r,t},\al_{r,t+1},\ldots,\al_{r,m},\al_{r+1,m},\ldots,\al_{l,m},\beta_1,\ldots,\beta_N), $$ where $\{\beta_1,\ldots,\beta_N\}$ are the elements in $\operatorname{supp}\bp$ such that $\beta_j> \al_{l,m}$. Since $\al_{r,t}\in \operatorname{supp}\tilde\bp$ we know: $$ \sum_{(j,k) \in \bp} s_{j,k}<\sum_{(j,k) \in \tilde \bp} s_{j,k}\le q_\bp^\lambda $$ and hence $\sum_{(j,k) \in \bp} s_{j,k}=\sum_{(j,k) \in \bp} s'_{j,k}\le q_\bp^\lambda-1=q_\bp^{\lambda-\om_i}$. \qed\\ For $\bs \in S(\lambda)$ we define a \textit{mutation} of $\bs$ as follows: \begin{definition} Let $$\beta = \sum_{(j,k) \in R} s_{j,k} \alpha_{j,k}$$ and suppose $$\beta = \sum_{(j,k) \in R} t_{j,k} \alpha_{j,k}$$ where $$t_{j,k} = 0 \text{ if } (j,k) \notin R_{\lambda} \; ; \; t_{j,k} \geq 0 \text{ if } (j,k) \in R_{\lambda},$$ for some $\bt = (t_{j,k}) \notin S(\lambda)$. Then we call $\bt$ a mutation of $\bs$. \end{definition} \begin{exam}\rm Let ${\mathfrak g} = \mathfrak{sl}_3$ and $\lambda = \omega_2$. Define $$\bs \text{ by } s_{1,3} = 1, s_{2,2} = 1 \text{ and } s_{i,j} = 0 \text{ else, }$$ and $$\bt \text{ by } t_{1,2} = 1, t_{2,3} = 1 \text{ and } t_{i,j} = 0 \text{ else.}$$ Then $\bt$ is a mutation of $\bs$. \end{exam} \begin{prop}\label{mutation} For $\bs \in S(\lambda)$ let $M^{\bs}_i$ be the minimal set. If $\bt^1$ is a mutation of $\bm^{\bs}_i$, $\bt = \bt^2 + \bt^1 \in S(\lambda)$ and $t^{1}_{j,k} \geq 0$, then $\bm^{\bs}_i \ll\bm^{\bt}_i$. \end{prop} \proof Recall (see Remark~\ref{minimalelements}) that $M^{\bs}_i=\{ (j_l, k_l) \; | \; l=1, \ldots, m \}$ with $$ 1\le j_1 < \dots < j_m\le i\le k_m < \dots < k_1. $$ Let $\bt^1$ be a mutation of $\bm^{\bs}_i$, so $t^1_{j,k}= 0$ for $(j,k) \notin R_i$. Then there exists $\sigma \in S_{m} \setminus\{ id \}$ such that if $t^1_{p,q} \neq 0$, then $(p,q) = (j_l, k_{\sigma(l)})$ for some $1 \leq l \leq m$. We can even assume that $\sigma(l) \neq l$ for all $l$, because otherwise $(j_l, k_l)$ is not mutated and appears in $\bm^{\bs}_i$ and $\bt^1$.\\ It is clear that $\bm^{\bt^1}_i \ll \bm^{\bt}_i$ (or equal), so it suffices to show that $\bm^{\bs}_i \ll \bm^{\bt_1}_i$.\\ Let $x = \sigma^{-1}(m)$, we claim that $M^{\bt}_i \subset \{ (j_1, k_{\sigma(1)}), \ldots, (j_x, k_{\sigma(x)}) \}$. Let $l > x$, then $j_x < j_l$ and $k_m > k_{\sigma(l)}$ (since $\sigma(l) \neq m$). So $(j_x, k_m) < (j_l, k_{\sigma(l)})$ for all $l > x$. \qed \begin{thm}\label{numberofelements} Let $\lambda = \sum_j m_j \omega_j \in P^+$. For each $\bs\in S(\la)$ fix an arbitrary order of factors $f_\al$ in the product $\prod_{\al>0} f_\al^{s_\al}$. Let $f^\bs=\prod_{\al>0} f_\al^{s_\al}$ be the ordered product in $U(\n^-)$. Then the elements $f^\bs v_\la$, $\bs\in S(\la)$, form a basis of $V(\la)$. \end{thm} \proof We will prove the claim by induction on $m = \sum_{j=1}^n m_j$. By Theorem~\ref{linearindipencetheorem} we know that the $f^\bs v_{\lambda}$ span the representation $V(\lambda)$, so $\dim V(\lambda)\ge \sharp S(\lambda)$. For the initial step $m=1$ the description of $S(\om_i)$ in Remark~\ref{minimalelements} shows that the tuples have all different weights and hence the $f^\bs v_{\om_i}$ are also linearly independent, which proves the claim for the fundamental representations. We assume that the claim holds for $\lambda$, we want to prove it for $\lambda + \omega_i$. We may assume again that $i$ is minimal such that $m_i \neq 0$. The highest weight vector $v_{\lambda} \otimes v_{\omega_i}$ generates $V(\lambda + \omega_i) \subset V(\lambda) \otimes V(\omega_i)$. We assume in the following that the roots are ordered in such a way that the $f_{\alpha}$ with $\alpha \in R_i$ are at the beginning. Every element $\bs \in S(\lambda + \omega_i)$ defines a vector of $f^{\bs} (v_{\lambda} \otimes v_{\omega_i}) \in V(\lambda + \omega_i)$. We want to show that these vectors are linearly independent, so we have to show \begin{equation} \label{lindepend} \sum_{\bs \in S(\lambda + \omega_i)} a_{\bs} f^{\bs}(v_{\lambda} \otimes v_{\omega_i}) = 0 \Rightarrow a_{\bs} = 0 \; \forall \; \bs \in S(\lambda + \omega_i). \end{equation} We may assume without loss of generality that all $\bs$ have the same weight, say $\bs\in S(\lambda + \omega_i)^\mu$. By Proposition~\ref{minimalset} we can split an element in $S(\lambda+\omega_i)$ such that $\bs = \bs_2 + \bm^{\bs}_i$, where $\bs_2 \in S(\lambda)$. Assume that we have a non-trivial linear dependence relation in \eqref{lindepend}. Fix $\bbs \in S(\lambda+ \omega_i)^\mu$ such that $a_{\bbs}\not=0$ in this relation and $a_{\bt} = 0$ for all $\bt$ such that $\bm^{\bbs}_i \ll \bm^{\bt}_i$. Consider first $\bbs = \bbs_2 + \bm^{\bbs}_i$, so we have \begin{equation} \label{rest} f^{\bbs}(v_{\lambda} \otimes v_{\omega_i}) = c_{\bm^{\bbs}_i} f^{\bbs_2}v_{\lambda} \otimes f^{\bm^{\bbs}_i}v_{\omega_i} + \text{\it other terms, } \end{equation} where $c_{\bm^{\bbs}_i}$ is a nonzero constant (product of binomial coefficients). All the terms occurring in the linear dependence relation \eqref{lindepend} can be rewritten as sums of terms of the form $f^{\br_2}v_{\lambda} \otimes f^{\br_1}v_{\omega_i}$. So in order to prove that necessarily $a_\bs=0$ for all terms in \eqref{lindepend}, it is sufficient to show that that the terms $f^{\br_2}v_{\lambda} \otimes f^{\br_1}v_{\omega_i}$ satisfying $wt(\br_2)=wt(\bbs_2)$ and $wt(\br_1)=wt(\bm^{\bbs}_i)$ are linearly independent. Let us first consider the possible terms in \eqref{rest} occurring among the \textit{other terms}. It is a sum of elements $f^{\br_2}v_{\lambda} \otimes f^{\br_1}v_{\omega_i}$, where $\br_2 + \br_1 = \bbs$ and $\br_1 \neq \bm^{\bbs}_i$. If $\operatorname{wt}(\br_1) = \operatorname{wt}(\bm^{\bbs}_i)$, then either $\br_1\in S(\omega_i)$, but then $\br_1=\bm^{\bbs}_i$ for weight reasons, or $\br_1\not\in S(\om_i)$. In the latter case the entries in $\br_1$ are zero for all $\al_{k,\ell}\not\in R_i$ because of the special choice of the ordering, and hence $\br_1$ has to be a mutation of $\bm^{\bbs}_i$. Then by Proposition~\ref{mutation}, $\bm^{\bbs}_i \ll \bm^{\br_1 + \br_2}_i = \bm^{\bbs}_i$ which is a contradiction. So the \textit{other terms} consist only of tensors of the form $f^{\br_2}v_{\lambda} \otimes f^{\br_1}v_{\omega_i}$, where $\operatorname{wt}({\br_2}) \neq \operatorname{wt}({\bbs_2})$ and $\operatorname{wt}({\br_1}) \neq \operatorname{wt}({\bm^{\bbs}_i})$, hence for proving linear independence we can neglect these terms. To obtain a non-trivial linear combination such that $a_\bt \neq 0$ for some $\bt\not=\bbs$, one needs an element $\bt \in S(\lambda + \omega_i)^\mu$ which can be splitted $\bt = \bt_2 + \bt_1$ such that $\operatorname{wt}({\bt_2}) = \operatorname{wt}({\bbs_2})$, $\operatorname{wt}({\bt_1}) = \operatorname{wt}({\bm^{\bbs}_i})$, and $f^{\bt_2} v_{\lambda} \neq 0, f^{\bt_1} v_{\omega_i} \neq 0$. Suppose that one has such a $\bt = \bt_2 +\bt_1 $ and $\bt_1 \notin S(\omega_i)$. By the same arguments as above, $\bt_1$ is a mutation of $\bm^{\bbs}_i$ and hence by Proposition~\ref{mutation}, $\bm^{\bbs}_i \ll \bm^{\bt}_i$. But in this case we have by assumption $a_{\bt} = 0$, contradicting the fact $a_\bt \neq 0$. It follows $\bt_1 \in S(\omega_i)$ and hence, by weight arguments, $\bt_1=\bm^{\bbs}_i$ and $\bt = \bt_2 + \bm^{\bs}_i$, where $\bt_2 \neq \bs_2$. So if a term of the form $f^{\bt_2}v_{\lambda} \otimes f^{\bt_1}v_{\omega_i}$ $\operatorname{wt}({\bt_2}) = \operatorname{wt}({\bbs_2})$, $\operatorname{wt}({\bt_1}) = \operatorname{wt}({\bm^{\bbs}_i})$ occurs in the linear dependence relation \eqref{lindepend}, then necessarily $\bt_1=\bm^{\bbs}_i$. Hence, by Proposition~\ref{minimalset}, $\bt_2 \in S(\lambda)$. Since the possible $\bt_2$ are different from $\bbs$ and by induction the terms $\{f^{\bt_2}v_{\lambda} \otimes f^{\bm^{\bbs}_i}v_{\omega_i}\mid\bt_2\in S(\lambda)\}$ are linearly independent, it follows $a_{\bbs} = 0$, contradicting the assumption $a_{\bbs} \neq 0$. Summarizing, we have shown that for the order fixed at the beginning of the proof the $f^\bs v_{\lambda+\om_i}$, $\bs\in S(\lambda+\om_i)$, are linearly independent and form a basis. This implies in particular that $\sharp S(\lambda+\om_i)=\dim V(\lambda+\om_i)$. Now by Theorem~\ref{linearindipencetheorem} we know that the $f^\bs v_{\lambda+\om_i}$, $\bs\in S(\lambda+\om_i)$, span $V(\lambda+\om_i)$ for any chosen total order. So, for dimension reason, they also have to be linearly independent for any chosen order. \qed \section{Proof of Theorem A and applications} In this section we collect some immediate consequences of the constructions in Sections 2 and 3. The proof of Theorem~\ref{numberofelements} shows: \begin{cor} \par\noindent $$ \dim V(\la) =\# S(\la)=\,\text{number of integral points in the polytope}\, P(\lambda). $$ \end{cor} By the defining inequalities (see \ref{polytopeequation}) for the polytope $P(\lambda)$ it is obvious that for two dominant integral weights $\lambda,\mu$ we have $P(\lambda)+P(\mu)\subseteq P(\lambda+\mu)$, and hence for the integral points we have $S(\lambda)+S(\mu)\subseteq S(\lambda+\mu)$, too. In fact, the reverse implication is also true: \begin{prop} $S(\la)+S(\mu)=S(\lambda+\mu)$. \end{prop} \proof Set $\nu=\lambda+\mu$ and write $\nu=\sum k_i\om_i$ as a sum of fundamental weights. Proposition~\ref{minimalset} provides an inductive procedure to write an element $\bs$ in $S(\nu)$ as a sum $\bs=\sum_{i=1}^n\sum_{j=1}^{k_i} \bm_{i,j}$ such that $\bm_{i,j}\in S(\omega_i)$ for all $1\le i\le n$, $1\le j\le k_i$. This sum can be reordered in such a way that $\bs=\bs^1+\bs^2$, $\bs^1\in S(\lambda)$, $\bs^2\in S(\mu)$, so $\bs\in S(\lambda)+S(\mu)$. \qed \bigskip As an interesting application we obtain a combinatorial character formula for the representation $V(\la)$. Let $P$ be the weight lattice and for $\bs \in S(\lambda)$ define the weight $$ \operatorname{wt}(\bs) := \sum_{1 \leq j \leq k \leq n} s_{j,k} \alpha_{j,k}. $$ Let $S(\lambda)^\mu$ be the subset of elements such that $\mu=\lambda-\operatorname{wt}(\bs)$ and let ${\tt S}(\lambda)^\mu:=\#\{ \bs \in S(\la) \; | \; \mu=\la-\operatorname{wt}(\bs)\}$ be the number of elements of this set. We obtain as a consequence of Theorem~\ref{basis}: \begin{prop} $$ char V(\la) =\sum_{\mu\in P}{\tt S}(\lambda)_\mu e^\mu. $$ \end{prop} \bigskip The big advantage of our approach is that it provides also a combinatorial formula for the graded character. Recall that $gr V(\la)$ carries an additional grading on each weight space $V(\la)^\mu$ of $V(\la)$: $$ gr V(\la)^\mu=\bigoplus_{s\ge 0} gr_s V(\la)^\mu=\bigoplus_{s\ge 0} V(\la)^\mu_s/V(\la)^\mu_{s-1}. $$ The graded character of the weight space is the polynomial $$ p_{\lambda,\mu}(q):=\sum_{s\ge 0}(\dim V(\la)^\mu_s/V(\la)^\mu_{s-1})q^s $$ and the graded character of $V(\lambda)$ is $$ char_q (V(\la))= \sum_{\mu\in P}p_{\lambda,\mu}(q) e^\mu. $$ We have a natural notion of a {\it degree} for the multi-exponents: \begin{dfn} $$\deg(\bs) := \sum_{1 \leq j \leq k \leq n} s_{j,k}.$$ \end{dfn} As an immediate consequence of Theorem~\ref{basis} we get \begin{cor*} $p_{\lambda,\mu}(q)=\sum_{\bs\in S(\lambda)^\mu} q^{\deg \bs}$ and $$ char_q(V(\la)) = \sum_{\bs \in S(\la)} e^{\la-\operatorname{wt}(\bs)} q^{\deg(\bs)}. $$ \end{cor*} Finally, we note that the results of Sections \ref{span} and \ref{li} imply the description of the annihilating ideal $I(\la)$. \begin{thm}\label{idealtheorem} \begin{equation}\label{ideal} I(\la)=S(\n^-)\left(\U(\n^+)\circ \mathrm{span}\{f_\al^{(\la,\al)+1}, \al>0\}\right). \end{equation} \end{thm} \begin{proof} Since $f_\al^{(\la,\al)+1}v_\la=0$ in $V(\la)$ for all positive roots $\al$, the right hand side of \eqref{ideal} belongs to $I(\la)$. Section \ref{span} shows that the relations in the RHS of \eqref{ideal} are enough to rewrite any element of $gr V(\la)$ in terms of the basis element $f^\bs v_\la$, $\bs\in S(\la)$. This proves our theorem. \end{proof} \section*{Acknowledgements} The work of Evgeny Feigin was partially supported by the Russian President Grant MK-281.2009.1, the RFBR Grants 09-01-00058, 07-02-00799 and NSh-3472.2008.2, by Pierre Deligne fund based on his 2004 Balzan prize in mathematics and by Alexander von Humboldt Fellowship. The work of Ghislain Fourier was partially supported by the DFG project ``Kombinatorische Beschreibung von Macdonald und Kostka-Foulkes Polynomen``. The work of Peter Littelmann was partially supported by the priority program SPP 1388 of the German Science Foundation.
\section{INTRODUCTION} The U.S.~Naval Observatory (USNO) CCD Astrograph Catalog (UCAC) project began observations in the Southern Hemisphere at Cerro Tololo Interamerican Observatory (CTIO) in January 1998. In October 2000 the first U.S. Naval Observatory CCD Astrograph catalog (UCAC1) \citep{2000AJ....120.2131Z} was published covering about 80\% of the southern sky with positions and preliminary proper motions for over 27 million stars. The Astrograph was moved to the Naval Observatory Flagstaff Station (NOFS) in October 2001 to complete the Northern Hemisphere observing after 2/3 of the sky were completed from CTIO. The second USNO CCD Astrograph catalog (UCAC2) \citep{2004AJ....127.3043Z} was released in July 2003 with the same level of completeness as in UCAC1, but with improved reduction techniques, early epoch plates for improved proper motions and extended sky coverage. All sky coverage for UCAC observations were completed in October 2004. The third USNO CCD Astrograph Catalog (UCAC3) \citep{u3r} is the first all-sky data release in the UCAC series, containing about 100 million entries with a slightly fainter limiting magnitude than in previous versions. The magnitude range for UCAC3 is roughly 8.0 to 16.3 mag in the UCAC bandpass (579-642 nm, hereafter UCAC magnitude). A detailed introduction into UCAC3 with comparisons to other catalogs and warnings for the users are given in \citep{u3r}. Any user is also urged to read the extensive ``readme'' file provided with the DVD or on-line release. The UCAC3 is based on a complete re-reduction of the pixel data aiming at more completeness with the inclusion of double star fitting, problem case investigations and the slightly deeper limiting magnitude \citep{u3x}. The final positions are based on the Tycho-2 \citep{2000A&A...355L..27H} reference frame as in UCAC2. However, Two-Micron All Sky Survey (2MASS) \citep{2006AJ....131.1163S} residuals are used to probe for systematic errors in astrometric reductions as will be explained below. In UCAC1 and UCAC2 it was shown that the 4K CCD in the astrograph camera has a relatively poor charge transfer efficiency (CTE), leading to coma-like systematic errors in the uncorrected stellar positions. The effect is seen mainly in the $x$-axis (right ascension), which is the direction of the fast readout of charge, while the $y$-axis (declination) shows a much smaller effect. In UCAC1 a simple empirical approach was used to correct for this effect with the basic assumption that the effect was linear along the $x$-axis and no assumption for a dependency on magnitude. For UCAC2 the empirical approach was extended with a more complex model as a function of $x,y$, and instrumental magnitude derived from flip observations of calibration fields. Flip observations are obtained by observing the same field with the telescope on one side of the pier (east or west) then repeat with the telescope on the other side. Thus two images of the same area in the sky are obtained which are rotated by $180^{\circ}$ with respect to each other. In both previous UCAC catalogs sub-pixel phase and field distortion errors have also been investigated. For UCAC1 the pixel phase was modeled with an empirical function showing an amplitude on the order of 12 mas resembling a sine-function. For UCAC2 the pixel phase errors were investigated further and found to also be a function of the full width at half maximum (FWHM) of the stellar image profiles. The field distortion pattern was modeled in both previous reductions by binning reference star residuals from individual frames used in the reductions. The new pixel data reductions up to $x,y$ data are described in \citep{u3x}. This paper describes the reductions following the $x,y$ data up to the CCD-based mean RA, Dec positions. Early epoch data and procedures to derive proper motions are presented in the UCAC3 release paper \citep{u3r}, while details about an important part of the early epoch data, the Southern Proper Motion (SPM) data will be given elsewhere \citep{spm}. \section {INPUT DATA} For the astrometric reductions of UCAC3 as described in this paper, two sets of input data are needed: the $x,y$ data from selected CCD observations, and reference stars. Here two different reference star catalogs are used for different purposes. \subsection {CCD Frame Selection} Out of the about 278,000 UCAC frames ever taken, all applicable survey frames, calibration field, and minor planet exposures are selected for the reductions presented here. Poor quality frames are excluded. Quality criteria include limiting magnitude, internal fit precision of high S/N stellar images, and mean image elongation. About 15\% of the observations are qualified as ``poor", see also \citep{2000AJ....120.2131Z}. All frames which pass those quality criteria are included. Then the all-sky completeness from the selected data are examined and frames which almost meet the quality criteria are included as needed to provide a complete all-sky coverage. The final UCAC3 catalog contains mainly the survey frame data. The minor planet observations will be published separately, while the calibration field observations are only used in the first reduction steps to derive corrections to systematic errors. A summary of the CCD observations is given in Table~\ref{framest1}. The frames taken along the path of Pluto are included here from a collaboration with L.~Young (SWRI) for occultation predictions. A total of about 50 fields in the sky observed at about 10 to 30 different epochs with multiple exposures each time were used as astrometric calibrations. These typically low galactic latitude fields around ICRF targets, equatorial calibration fields, and open clusters were observed about 10 to 30 times during the project and with the telescope flipped between orientations east and west of the pier to provide a reversal of RA, Dec orientation in $x,y$ space. Data from these fields are utilized to derive certain systematic error corrections (see below). However, these frames are not included in the final UCAC3 reductions based on the Tycho-2 reference catalog. Frames around extragalactic link sources taken with the astrograph at times of deep CCD observations (mostly with the KPNO and CTIO 0.9 m telescopes) are not included here either. These data are still under investigation and a separate paper is in preparation regarding the optical link to ICRF quasars. \subsection {$x,y$ Data} For all astrometric reductions described below only the $x,y$ results from pixel data profile fit model five (symmetric Lorentz profile with pre-set $\alpha, \beta$ shape parameters) are used. This fit model has five free parameters per single star image (background, amplitude, center $x,y$, width of profile), similar to the more familiar two-dimensional Gauss model. However, the Lorentz profile matches the observed point-spread function (PSF) of our data significantly better than the Gaussian model. For more details and explicit PSF model functions see the UCAC3 pixel reduction paper \citep{u3x}. Double star fits of blended images are based on the same Lorentz image profile model, however three more free parameters are used for the two center coordinates and the amplitude, respectively, of the secondary component. Both the primary and secondary component are fit simultaneously. A single width of the profiles for both components of a double star and a single background level parameter is used in this fit. The $x,y$ data files also contain internal errors on the center coordinates derived from the least-square fits, two instrumental magnitudes based on the profile fit model and a real aperture photometry, respectively, and several auxiliary flags from the raw pixel reduction step. For more information about the raw data reductions see the UCAC3 pixel reduction paper \citep{u3x}. \subsection {Reference Stars} The 2MASS catalog does not provide proper motions. However, the UCAC and 2MASS observations were made almost at the same epoch and the 2MASS positions are of high quality with random errors of about 70 mas per coordinate for well exposed stars and small systematic errors \citep{jd16}. Due to the deep limiting magnitude and high density the 2MASS catalog is an excellent tool to probe UCAC data for systematic errors on a statistical basis, allowing to stack up many residuals as a function of a large number of parameters, as will be described below. Table~\ref{framest2} lists the number of available observations of reference stars, each providing a residual along RA ($x$) and Dec ($y$). For this purpose a subset of the 2MASS point source catalog is constructed. Stars are selected from the Naval Observatory Merged Astrometric Dataset (NOMAD) using the 2MASS identifier and imposing a limit of V or R $\le$ 16.5 to select 116,247,341 2MASS stars. For these, the original 2MASS positions are retrieved and matched with UCAC observations on a frame-by-frame basis. Proper motions of this reference star catalog are taken out of the NOMAD catalog. Even if these are not very accurate for faint stars, they bridge the few years of epoch difference to UCAC data to avoid large-scale biases from galactic dynamics. For most stars in the southern hemisphere, the epoch difference between 2MASS and UCAC is $\sim \pm$ 1 yr, while it gradually increases for stars at higher declinations. Near the north celestial pole the epoch difference is about four years. The 2MASS data are used only to derive most of the systematic error corrections of the UCAC $x,y$ data. As with UCAC2, the Tycho-2 catalog is used as the only source for reference stars in a final astrometric reduction to obtain UCAC3 positions on the International Celestial Reference Frame (ICRF). Only stars from the Tycho-2 catalog indicated as having good astrometry have been used for this reduction. Tycho-2 proper motions have been used to propagate the positions to the epoch of the individual UCAC $x,y$ observations. Both reference star catalogs are sorted by declination and stored in binary direct access files. For each catalog a match to the $x,y$ data are performed to produce cross-reference output files per CCD frame containing the record numbers from the $x,y$ direct access data and the reference star catalog, as well as the magnitude of the star, B$-$V color, and astrometry flag (for Tycho-2). This scheme allows a fast runtime for the many passes through all the data to perform the astrometric reductions without the need for a match of reference stars each time. \section {ASTROMETRIC REDUCTIONS} For the astrometric reductions in UCAC3 extensive systematic error investigations are performed. The largest systematic error is caused by the poor charge transfer efficiency (CTE) of the detector, followed by geometric field angle distortions. An iterative approach was adopted utilizing 2MASS residuals to determine each of these effects in turn as described below. The idea here is to use an empirical approach as done in the past, but investigate further dependencies to better model the effects seen in the 4K CCD pixel data. Preliminary results were presented earlier \citep{2009AAS...21347305F}. A pure magnitude equation is derived from internal calibration observations only. Finally the position shifts as function of the sub-pixel location of an image in the focal plane (the sub-pixel phase error) is derived, however, the $x,y$ data used for that are the original measures before applying the other corrections. Systematic position offsets as a function of color and differential color refraction due to Earth's atmosphere are small (about 5 mas) due to the narrow UCAC spectral bandpass. No such corrections are applied for the UCAC3 catalog. \subsection {Preliminary Field Distortion Pattern} In the first step of the astrometric reductions, the CTE caused systematic errors (coma-like) are approximated by a linear model as a function of magnitude and $x$ pixel coordinate, similar to the CTE modeling of UCAC1. This takes out about 80\% of the CTE effect. The pre-corrected $x,y$ data are then used in a preliminary astrometric reduction with 2MASS reference stars to produce an approximate field distortion pattern (FDP), i.e.~systematic errors purely based on the $x,y$ location of a stellar image on the area of the detector. Although the coma-like corrections for the CTE effect should not bias the purely geometric FDP corrections, the scatter in the data is largely reduced by correcting for most of the CTE effect first. This leads to a more accurate FDP as would be generated without applying any CTE corrections. Maximal systematic errors in the FDP are about 24 mas, with typical corrections in the 5 to 10 mas range. This preliminary FDP is then applied to the otherwise uncorrected $x,y$ data to analyze the CTE effect from scratch in the following step. The same reasoning applies here; with good FDP correction in hand the scatter of the residuals is reduced to accurately probe the systematic errors induced by the poor CTE. \subsection {CTE Effect} When the CCD reads out an image, the electrons are transferred from pixel$-$to$-$pixel until the charge reaches the output register. The low CTE of the 4K CCD causes charge to be left behind as this transfer occurs, leading to slightly asymmetric images. The amount of asymmetry and the derived $x,y$ center of stellar images with respect to the unaffected position depend on the $x$ pixel location, the brightness of the star, the length of the exposure, and other factors. As found here and in previous UCAC reductions the CTE effect is by far the most substantial systematic error seen in the raw UCAC $x,y$ data amounting up to about 200 mas. The effect is predominantly seen in the $x$-axis along right ascension and increases from no effect near $x = 0$ pixel to a maximum near $x = 4094$ pixel, as evident in Figure~\ref{CTEf1}. A similar effect is also seen in the $y$-axis, but to a lesser degree because of a slower charge transfer in the direction of declination. For the reductions, the UCAC frames are separated into individual data sets depending on observation site (CTIO or NOFS) and telescope orientation (east or west). Over 216,000 frames are used for the reductions, see Table~\ref{framest1}. Most frames at CTIO were observed with the telescope west of the pier, while at NOFS the regular observing was performed east of the pier. Because the instrument had to be disassembled for the move from CTIO to NOFS the camera alignment and other instrument properties are slightly different at both sites. This explains the slightly different patterns for systematic error corrections found for the two sites. However, the 2MASS reference star catalog is dense enough to provide a sufficient number of residuals for statistically significant results even on relatively small sub-sets of the UCAC data. For UCAC3 we found that the CTE systematics show a dependence on FWHM, brightness of the star, exposure time, location on the chip ($x,y$), camera orientation (east, west) and observing site (CTIO, NOFS). For example as Figures~\ref{CTEf1} and \ref{CTEf2} show the CTE caused different systematic position offsets for the $x$ coordinate as a function of magnitude for 20 and 150 sec exposures, respectively. The residuals with respect to the 2MASS reference stars are split into two major data sets for investigating the CTE effect, CTIO west and NOFS east. A plotting program is then used to display and determine empirical corrections to create a complex look-up table. The table is split up into four different FWHM bins keeping a roughly equal number of frames per bin, with half step magnitude bins ranging from 8 to 16 magnitude for all standard exposure times of 20$-$200 seconds duration, and 5 bins along the $x,y$ axes. Each data set is evaluated and look-up tables are created to correct for the residuals, (see sample, Table~\ref{CTEt1}). For each of the main data sets, CTIO west and NOFS east, tables are created separately for the $x$ and $y$ axes. After testing we found that these tables could also be used for the data of the other configurations, (CTIO east and NOFS west) corresponding to the observation site. After applying corrections and re-running the residuals the correction tables are continuously updated until the residuals flattened out. The largest correction from the residuals for CTIO is 204 mas in the $x$-axis and $-$32 mas in the $y$-axis. For NOFS the largest correction from the residuals is 216 mas in the $x$-axis and $-$67 mas in the $y$-axis. \subsection {Pure Magnitude Equation} The 2MASS positions are uncorrelated to the $x,y$ pixel coordinates of the UCAC observations. Thus any systematic errors seen in the 2MASS$-$UCAC residuals as a function of UCAC $x,y$ and also as a function of magnitude times these coordinates (coma terms) are inherent in the UCAC data and can be corrected with the above procedure. However, this is not the case for a pure magnitude dependent systematic error, which could originate in either or both catalogs. With the systematic error corrections applied to the UCAC $x,y$ data as described above any possible pure magnitude equation from 2MASS was transferred into the UCAC positions. Different catalogs have typically different magnitude equations. As an example Figure~\ref{TYCf1} shows the residuals as function of magnitude from a reduction of UCAC data with Tycho-2 reference stars, after applying all systematic error corrections based on the 2MASS reductions. This clearly shows a difference in magnitude equation between 2MASS (= UCAC system at this point) and Tycho-2 without knowing what the error free positions might be. The flip observations of calibration fields are used to determine the overall, pure magnitude dependent systematic errors in the UCAC data independent of any external reference star catalog. With these observations the same field in the sky has been observed with sets of frames rotated by 180$^{\circ}$ with respect to each other. Any $x,y$ coordinate offset as function of magnitude shows up in the residuals of a transformation of east versus west frames $x,y$ data. We derived overall magnitude equation slope terms for long and short exposure frames and both sites separately. These are then applied globally in the final astrometric reductions. After applying all corrections a small magnitude term of a few mas/mag is still seen in the residuals of the Tycho-2 reductions of UCAC data as shown in the UCAC3 release paper \citep{u3r}. This indicates such a magnitude equation is present in the Tycho-2 catalog itself, which is found to vary as a function of declination zone as one would expect. It would be preferable to derive all systematic error corrections from internal calibration observations. However, the flip observations alone do not allow us to do so because of the degeneracy between a pure magnitude equation and coma-like terms. Only after correcting the UCAC $x,y$ data for coordinate dependent (including coma) terms do the flip observations allow for a unique solution of the magnitude equation. Of course the assumption here is a constant magnitude equation, not changing from exposure to exposure. This assumption can not easily be made for photographic astrometry, which is affected by a highly non-linear detector, but should hold better for CCD data. \subsection {Field Distortions} Field distortion patterns (FDPs) are derived for UCAC3 by binning thousands of reference star residuals of individual CCD frames using the same procedure as in the previous UCAC reductions. These reductions are performed on the CTE corrected data but without applying the preliminary FDP used before. From deriving FDPs of various subsets of the data we found that the FDP is almost constant except for a small difference depending on observing site (CTIO or NOFS). Residuals are created using all good survey frames with respect to 2MASS reference stars split into CTIO west and NOFS east data sets. The data using opposite camera orientations than used for most of the observations (i.e.~CTIO east and NOFS west) did not have significantly different field distortions so only two correction tables are used for the final reductions. FDP corrections for frames taken with the camera orientation not used frequently are applied by rotating the FDP of the data with the large amount of observations at that site. Figure~\ref{FDPf1} (top) shows the FDP for CTIO west with vectors up to 23 mas in length. The FDP for NOFS east is slightly different with vectors up to 24 mas at some bins. In Figure~\ref{FDPf1} (bottom), we show the difference (CTIO west$-$NOFS east) of the two data sets. Although these differences are small (below 6 mas) they are systematic and well determined. Therefore we choose to use a separate FDP map to correct the data of each site. \subsection {Subpixel Phase Errors} After the above mentioned systematic errors have been corrected the 2MASS reference star catalog is used again to generate residuals from all applicable UCAC survey frames. Residuals are analyzed as a function of the original $x$ and $y$ pixel coordinate fraction (sub-pixel phase) before other corrections are applied. Various sub-sets of the data are looked at. Systematic errors are found to be a function of the FWHM of the image profiles. Figure~\ref{SUBPf1} gives some examples. The results are found to be independent of exposure time, as expected. However, a slight difference between the CTIO and NOFS data is found. The amplitude of the sub-pixel phase dependent systematic errors in the star positions is shown in Figure~\ref{AMPf1}, here for the corrections to the $x$ coordinate; those for the $y$ coordinate are somewhat smaller. All sub-pixel phase systematic corrections are smaller than what was found in UCAC2. This is a consequence of using an image profile model in UCAC3 (Lorentz profile) which better fits the true PSF than was the case for UCAC2 (Gauss model). However, the function of the sub-pixel phase systematic errors are more complex in the UCAC3 than UCAC2 data, where a simple sine and cosine term were sufficient. For the UCAC3 data we had to expand to three sine and three cosine terms in order to fit the sub-pixel phase errors sufficiently well (Figure~\ref{SUBPf1}). These six parameters are determined separately for 12 sets of data binned by FWHM (from 1.5 to 3.0 pixels), and split by NOFS and CTIO data. Calibration tables are then generated for equal steps along FWHM by interpolation and $x,y$ corrections applied, separately for each coordinate, based on these tables. \section {MEAN POSITIONS} Positions of all detected objects are obtained frame-by-frame from a final astrometric reduction with the Tycho-2 reference star catalog and correcting raw $x,y$ data first for the sub-pixel phase errors, then for systematic errors as a function of $x,y$ (FDP), then for mixed terms of coordinate with magnitude (CTE effect), and finally for a pure magnitude equation, as explained above. Apparent places and refraction are corrected rigorously using the Software for Analyzing Astrometric CCD (SAAC) code \citep{saac}, which also utilizes the Naval Observatory Vector Astrometry Subroutines (NOVAS) code \citep{novas}\footnote{$http://aa.usno.navy.mil/software/novas/novas\_info.php$}. The thus obtained positions are on the ICRF at individual epoch of each CCD frame (between 1998 and 2004) and are output to FPOS (final position) files. Previously identified and flagged observations of minor planets and high proper motion stars are output to separate files. All other individual positions are output by declination zones and then sorted by declination. Weighted mean positions are calculated from the individual images of each star, generating a running star number, MPOS (mean position file) on the fly. Over 139 million objects are identified at this step. All MPOS entries are then matched with early epoch star catalogs and another, more comprehensive 2MASS extract containing about 338 million objects. These 2MASS stars are selected directly from the 2MASS point source catalog without going through NOMAD. The R magnitude was estimated based on the 2MASS near-infrared J$-$K color and stars with R $\le$ 17.0 or J $\le$ 15.5 are selected. The unique identifier for stars matched across catalogs is the MPOS star number. Individual early epoch positions are output together with MPOS entries (CCD epoch observations) and sorted by MPOS number. Weighted proper motions and mean positions are then calculated to obtain the UCAC3 release catalog data \citep{u3r}. Objects which did not have either a reasonable proper motion determined or could not be matched with 2MASS are dropped at this point. Only these compiled catalog mean positions and proper motions are published in UCAC3, not the MPOS or FPOS data, which likely will be made available for the final UCAC4 release after further updates. \section{COMPARISON WITH HIPPARCOS} A total of 1510 Hipparcos stars in the 8 to 12 magnitude range are randomly selected (about 300 all sky per magnitude interval) and flagged in the UCAC data. These stars are not used as reference stars in test reductions using Tycho-2 reference stars. Thus the obtained positions are field star positions from UCAC observations independent of the Hipparcos and Tycho catalog positions. Individual UCAC observed positions are then compared to the original \citep{hiporig} and new Hipparcos reductions \citep{2007ASSL..350.....V} at the epoch of UCAC observations, using Hipparcos mean positions, proper motions and parallaxes. After excluding outliers ($\ge$ 200 mas position difference in either coordinate), RMS values over observations in each bin are calculated (Table~\ref{TYCt1}). Similarly sorting all observations by magnitude or color, respectively and binning over 100 observations lead to the plots shown in Figure~\ref{UHipmag} and \ref{UHipcol}. The expected position errors from UCAC3 and Hipparcos data at the epoch of our UCAC observations are also presented in Table~\ref{TYCt1} together with the expected RMS of the combined error and the ratio of expected to observed scatter, separately for each coordinate. In all cases the observed errors (from the scatter of the UCAC3$-$Hipparcos position differences) is slightly smaller than the expected errors as calculated from the combined formal errors for the same observations, thus at least some of these are overestimated (see also discussion section below). UCAC position differences of those sampled Hipparcos stars do not show systematic errors as a function of magnitude or color exceeding about 10 mas over the range sampled. Plots with respect to the original or new Hipparcos catalog are almost identical. However, 23 Hipparcos stars (1.5\% of this sample) show very large differences (between 300 and 600 mas in either coordinate) when comparing the new reduction Hipparcos positions with the UCAC3 positions at UCAC epoch. A similar number of stars is found when comparing with the original Hipparcos Catalogue; however, for not exactly the same stars. All possible combinations of inconsistencies between the two Hipparcos solutions and UCAC data are found, with two of the three positions or all three separated by several standard errors of their internal errors. \section {DISCUSSION} The use of an image profile model better matching the actual PSF than a Gaussian model is essential for the astrometric reductions of blended images. A better matching model also does reduce the amplitude of the sub-pixel phase error and is advisable to be used when no such corrections are being applied to the data. In particular, a comparison of Figure~\ref{AMPf1} with a similar figure of the UCAC2 paper show that with a Gaussian model and 2.0 pixel/FWHM sampling the pixel phase error has an amplitude of 11 mas, while with the image profile model 5 as used in UCAC3 this amplitude is only about 6 mas. The use of such a PSF profile model (still with the same number of fit parameters per star as the traditional Gaussian model) allows to neglect positional errors as function of sub-pixel phase completely for a sampling of about 2.5 pixel/FWHM or larger without the need to investigate possible other dependencies of this systematic error as a function of other things. However, for single stars and with calibration data in hand to correct for the position offsets caused by a sub-pixel phase dependency, the use of a more sophisticated image profile model than a Gaussian might not have an apparent advantage for astrometric reductions. The slight difference in amplitude of the sub-pixel phase corrections between CTIO and NOFS data is surprising. It could be caused by a slightly different, observed PSF between the two sites, even for the same seeing (FWHM). Whether this is caused by differences in the instrument, guiding or atmosphere is currently not known. The small systematic errors of UCAC based positions of randomly selected Hipparcos stars confirms the good correction of UCAC3 epoch positions as function of magnitude and color, at least for the 8 to 12 magnitude range. With UCAC3 positions agreeing with Hipparcos data the magnitude equation seen in residuals with respect to Tycho-2 is an indication for such small systematic errors in the Tycho-2 catalog. These are likely introduced through the proper motions, thus the early epoch, ground-based data, as also indicated in the UCAC3 release paper. The random errors in the observed UCAC$-$Hipparcos position differences are even slightly smaller than the expected, combined formal errors. For the new Hipparcos reductions the difference is only a few percent, while for the comparison with the original Hipparcos Catalogue data the observed errors are about 10\% smaller than expected. This indicates a slightly overestimated error in the original Hipparcos Catalogue proper motions. The formal position errors for the individual UCAC observations do include the formal image profile fit error, and the conventional plate adjustment error propagation. The weighting scheme used in this individual CCD frame least-square adjustments also include an estimated error contribution from the turbulence in the atmosphere, scaled by the exposure time. The mismatch between the actual PSF and the image profile model can lead to an overestimation of the center position errors, particularly for stars as bright as this sample of Hipparcos stars, which would explain the slightly smaller than expected scatter in the Hipparcos to UCAC position differences. The exclusion of outliers at an arbitrary limit of 200 mas could be another possible explanation. At the faint end of Hipparcos (11th magnitude) the Hipparcos catalog positions are of comparable precision to typical mean UCAC positions (based on 4 images) at their about 2000 epoch. The next step after UCAC, the USNO Robotic Astrometric Telescope (URAT) program \citep{urat} to begin in 2010 thus will likely be capable of improving proper motions of individual Hipparcos stars significantly. \acknowledgments The entire UCAC team is thanked for making this all-sky survey a reality. For more detailed information about ``who is who" in the UCAC project the reader is referred to the readme file and UCAC3 release paper. The California Institute of Technology is acknowledged for the {\em pgplot} software. More information about this project is available at \\ \url{http://www.usno.navy.mil/usno/astrometry/}.
\section{Introduction} Historically, prepositions have not enjoyed the attention of nouns and verbs, being until recently relegated to the status of {}``an annoying little surface peculiarity'' \cite{jackendoff}. However, linguistics tells us that different syntactic categories contain distinct semantic characteristics which are often exclusive to members of that category \cite{levin91}. This raises two distinct yet equally important questions: How are such categories found syntactically, and how are their characteristics expressed semantically? The famed linguist Sir Randolph Quirk states that {}``a preposition expresses a relation between two entities, one being that represented by the prepositional complement'' \cite{quirk85}. In this paper we describe the development of a heuristics-based system. \section{Theoretical Foundations} Through a survey of the literature, it is clear that the study of prepositions is becoming more intricate, and is segmented into a few focussed areas. Our work is therefore motivated towards the construction of a sequential system of increasingly complex levels, each of which is representative of one of these focussed areas. The system is organized in such a way that the product of one level becomes a dependency of the next as outlined by the following sequence: \begin{lyxlist}{00.00.0000} \item [{Level~0:}] From Part-of-Speech annotated text, minimal prepositional phrases are found at the syntactic level according to a context-free grammar (CFG), and not categorized. \item [{Level~1:}] Minimal prepositional phrases are augmented with a set of labels indicating classes of semantic roles, by means of rule-based heuristics. \item [{Level~2:}] The proper attachment of the prepositional phrase is attempted with shallow heuristics based on results of Levels 0 and 1, in case of ambiguity. \item [{Level~3:}] Semantic characteristics of the PP and its co-predicate phrases are analyzed in order to perform attachment 'intelligently', and do discover more thorough semantic relations. \end{lyxlist} Each level is described in more detail in Sections \ref{sub:Discovery-of-Minimal} through \ref{sub:Consideration-of-Semantics}. \subsection{\label{sub:Discovery-of-Minimal}Discovery of Minimal PPs - Syntactic Analysis} The first stage is to locate prepositional phrases within a text grammatically, irregardless of semantic function. For this we turn to linguistics, which studies PPs by their production and by their expansion as outlined below. \subsubsection{Production of PPs} Prepositional phrases generally attach to noun phrases and verb phrases, usually acting adjectively for the former and adverbially for the latter \cite{dickphrases}, as in the example \begin{verse} {\footnotesize (S}\\ {\footnotesize ~~~(NP }\\ {\footnotesize ~~~~~~(DET the )~~(HEAD boy) }\\ {\footnotesize ~~~~~~(}\textbf{\footnotesize PP}{\footnotesize{} }\\ {\footnotesize ~~~~~~~~~in }\\ {\footnotesize ~~~~~~~~~(NP}\\ {\footnotesize ~~~~~~~~~~~~(DET the )~~(HEAD shop) }\\ {\footnotesize ~~~~~~~~~)))}\\ {\footnotesize ~~~(VP}\\ {\footnotesize ~~~~~~is waiting}\\ {\footnotesize ~~~~~~(}\textbf{\footnotesize PP}{\footnotesize }\\ {\footnotesize ~~~~~~~~~at}\\ {\footnotesize ~~~~~~~~~(NP}\\ {\footnotesize ~~~~~~~~~~~~(DET the)~~(HEAD corner)}\\ {\footnotesize ~~~~~~~~~)))}\\ {\footnotesize )}{\footnotesize \par} \end{verse} At first glance, we should be able to augment grammatical noun and verb phrases by simply adding rules of the form $NP\Leftarrow NP\, PP$ and $VP\Leftarrow VP\, PP$ \cite{jurafsky}. However this can lead to overgeneration, especially in cases where there exist ambiguity with regards to the part-of-speech of a potential preposition. For example, in {}``\texttt{\footnotesize Turn off/RB the light}'', {}``\texttt{\footnotesize off}'' is used as an adverb, but if we are not careful with our expansion rules for $VP$, we could easily generate {}``\texttt{\footnotesize (VP turn (PP off/IN (NP the light)))}'' if the tagging is done incorrectly. Some such errors in tagging cannot be avoided within the scope of this project, but careful augmentation of existing rules, such as in $NP\Leftarrow DET\, MOD\, HEAD\, PP$, can result in a more fine-tuned grammar, as will be shown in our results (see Section and $VP$ expansions producing $PPs$are given in Appendix B. \subsubsection{Expansion of PPs} Though there is some debate as to the semantic function played by the prepositional phrase \cite{francez2003}, the syntactic nature of a prepositional phraseis relatively universally accepted. It is very safe to define the prepositional phrase specifically as havinga preposition as its head, followed by a noun phrase or an entity whichis always the direct object \cite{dickphrases}. A full listing of rules for $PP$ expansions are given in Appendix A and B. \subsubsection{Non-minimal PPs} We will often encounter sequences of contiguous prepositional phrases, as in {}``\texttt{\footnotesize I saw the fool}{\footnotesize{} }\texttt{\textbf{\footnotesize on/IN}}{\footnotesize{} }\texttt{\footnotesize the hill}{\footnotesize{} }\texttt{\textbf{\footnotesize with/IN}}{\footnotesize{} }\texttt{\footnotesize the telescope}''. In such circumstances, we are principally concerned with PP-attachment, which is a significant grammatical challenge, but which also involves semantic analysis, hence its discussion is delayed until Section \ref{sub:PP-Attachment}. Suffice it to say, we will with some regularity encounter such a sequence of PPs where each PP modifies the same predicative, and hence we add the production rule $PP\Leftarrow PP\, PP$ to deal with such a circumstance. \subsubsection{The Mechanism of Discovery} The mechanism for discovering prepositional phrases is an implementation of the Earley parser in Scheme. While this provides a simple interface to grammatical tree-expansions, based purely on syntactic context-free rules -- the algorithm is devoid of any stochastic or world knowledge, and hence cannot be used for more complex grammatical or semantic analysis. The instantiated grammar is an instance of a partial parser PP-chunker which searches exclusively for prepositional phrases. \subsection{Semantic Role Annotations \& Categorization} Prepositions convey significant semantic relations in text, and provide {}``the principal means of conveying semantic roles for the supporting entities referred to in a predication'' \cite{techreport}. They are, however, highly ambiguous - with closely related word-senses and a relatively high degree of internal polysemy. Furthermore, definitions as to \emph{how} semantic roles are conveyed, and \emph{how} supporting entities relate are varied and often imprecise. At present, it is not feasible for automatic systems to attain a degree of semantic granularity comparable to that of collegiate dictionaries, but we can at least categorize different uses of the prepositional phrase by broad semantic characteristics. \subsubsection{\label{sub:Semantic-Role-Annotations}Semantic Role Annotations} The Penn Treebank described the prepositonal phrase as having semantic-role subcategorizations defined by case-style relations, the 7 most frequent of which are shown with their associated frequencies of occurrence in Table \ref{cap:Subcategorization-of-augmented}. Although by no means a thorough description of the semantic relations within a text, such subcategorizations allow for a suitable indication of the manner in which a prepositional phrase is used, and therefore to how they modify the semantics of the phrase to which they attach. \begin{table}[H] \begin{centering} \begin{tabular}{|c|c|c|} \hline tag & Freq. & Description\tabularnewline \hline \hline PP-LOC & 17220 & locative\tabularnewline \hline PP-TMP & 10572 & temporal\tabularnewline \hline PP-DIR & 5453 & direction\tabularnewline \hline PP-MNR & 1811 & manner\tabularnewline \hline PP-PRP & 1096 & purpose\tabularnewline \hline PP-EXT & 280 & extent\tabularnewline \hline PP-BNF & 44 & beneficiary\tabularnewline \hline \end{tabular} \par\end{centering} \caption{\label{cap:Subcategorization-of-augmented}Subcategorization of augmented PPs, ordered by frequency of occurence in the \cite{treebank}.} \end{table} These semantic relations can be attached to any verb complement, but more frequently occur with noun phrases and their clauses. The University of Pennsylvania provides online annotated texts from SIGLEX'99 \cite{seize} \cite{trump} with parsed examples showing each of these categories: \begin{lyxlist}{00.00.0000} \item [{PP-LOC:}] {}``...a federal grand jury \textbf{(PP-LOC in (NP Newark))}...'' \cite{seize} \item [{PP-TMP:}] {}``The people who suffer \textbf{(PP-TMP in (NP the short term))}...'' \cite{seize} \item [{PP-DIR:}] {}``Citicorp had discussed lowering the offer \textbf{(PP-DIR to (NP \$250 a share))}'' \cite{trump} \item [{PP-MNR:}] {}``He could be left \textbf{(PP-MNR without (NP top-flight legal representation))}...'' \cite{seize} \item [{PP-PRP:}] {}``...prosecutors told Mr. Antar's lawyers that \textbf{(PP-PRP because of (NP the recent Supreme Court rulings))}...'' \cite{seize} \item [{PP-EXT:}] {}``AMR declined \textbf{(PP-EXT by (NP \$22.125))}...'' \cite{trump} \item [{PP-BNF:}] {}``I baked a cake \textbf{(PP-BNF for (NP Doug))}'' \cite{treebank} \end{lyxlist} It is with this simple framework of 7 semantic annotations that we choose to develop our system. Although our heuristics will be designed to suit these categories, alternative classifications of prepositional phrases can be used to help obtain insight towards this process. \subsubsection{Alternative classifications} Alternative classifications for prepositional phrases have been discussed within the domain of semantic encoding in lexica \cite{eagles96} where labels are assigned to prepositions but describe whole phrases according to whether they apply to PPs modifying predicative heads (verbs and predicative nouns) or to PPs modifying nonpredicative heads (nonpredicative nouns). This distinction is justified semantically and syntactically% \footnote{From a syntactic point of view, the choice of the prepositional label modifying nonpredicative heads is much more restricted than for predicative heads, although more predominantly in the romance languages than in the germanic family. Generally, nonpredicative prepositional modifiers 'behave' more like adjectives, whereas predicative prepositional modifiers 'behave' more like adverbs. \cite{seize}% }, and most labels associated with predicative-head modifiers coincides directly with the semantic role annotations in Section \ref{sub:Semantic-Role-Annotations}. The subtlety such a distinction makes is shown for modifiers of non-predicative heads, all of which describe quality modifications. Consider, for instance, the difference between a place-position (locative) modifier in {}``the report is ON THE TABLE'' and a quality-place modifiers in {}``oranges FROM SPAIN'' (origin) and {}``a road THROUGH EGYPT'' (path). Even subtle semantic differences in positional modifiers can be noted depending on the predication of the head. Although such a distinction is not replicated in our system, the motivation behind the analysis of the semantic qualities of the head is applied theoretically to our heuristics. \subsubsection{Temporal PPs (TPPs)} Of those semantic categorizations common across multiple paradigms, temporal prepositional phrases have received special attention. Ian Pratt et al. \cite{pratt97}, show that temporal prepositional phrases with NP complements, and also with PP complements (validating our non-minimal grammar decision!), and even sentential complements with quantification-restrictions and structural ambiguities can be grouped into a unified theory of \emph{generalized temporal quantifiers} which serve as meanings for temporal PPs specifically, but also to NPs and sentences. Their theory differs from the normal categorization theory in that their semantics rely heavily on the $\lambda$-calculus and warp functions, and hence is only used as a theoretical landmark. It should be noted that in response to Pratt et. al, Nissim Frances et al. \cite{francez2003}% \footnote{Who, ironically, contributed to the paper under criticism.% }, argues that the unorthodox semantic operations and lack of syntactic cues in that paper makes such an approach infeasible for practical purposes, and that a simpler syntactic framework (\emph{indexical prepositional phrases}), would better be employed, which is an approach somewhat closer to our implementation. \subsubsection{\label{sub:Categorization-Heuristics}Categorization Heuristics} The heuristic approach to the problem of semantic categorization is easily implemented in a transformational PERL script that accepts simple PP partial parses and transforms the tag \textbf{PP} to \textbf{PP-XXX} (where XXX is one of the annotations of Section \ref{sub:Semantic-Role-Annotations}). Syntactic cues from the partial parse include the following: \begin{enumerate} \item The lexical entry of the head preposition \item The head of the component noun phrase \end{enumerate} It is tempting to produce heuristics based exclusively on the preposition alone (1), since intuitively the preposition should indicate the manner in which a PP is being used. For instance, in the examples {}``\texttt{\scriptsize Put the letter IN the mailbox}'', {}``\texttt{\scriptsize the dog is IN the space capsule}'' and {}``\texttt{\scriptsize The farmer IN the dell}'', the preposition {}``in'' is always used as a locative preposition. A moment's reflection will reveal that this approach is \emph{extremely} naive, because \emph{most} prepositions can be used for multiple functions, as in the following examples: \begin{itemize} \item ex. {}``The transient sleeps WITHIN the cardboard box'' (LOC) vs. {}``I'll be back WITHIN three weeks'' (TMP) \item ex. {}``I kicked the ball TOWARDS the net'' (DIR) vs. {}``My politics tend TOWARDS the left'' (MNR) \item ex. {}``I'll see you IN a couple of weeks'' (TMP) \end{itemize} Our heuristics cannot in the vast majority of cases consider the preposition alone because prepositions can often be used under different semantic relations. However, some prepositions, such as {}``when'' or {}``because'' are so often associated with a single semantic function (temporal and purpose, respectively), that we \emph{can} make minimal use of the preposition's lexical entry, but not exclusively. We implement a two-layer system of subsuming heuristics - the first which makes a {}``guess'' at the class depending on lexical knowledge associating words with semantic classes. For instance, \begin{lyxlist}{00.00.0000} \item [{TMP:}] when, until, in, during, after, before, while, under, over, then, since, around, at, throughout \end{lyxlist} This will of course lead to instances where certain ambiguous prepositions such as {}``for'' are always guessed to be of the most frequent category - but this is excusable because of the second heuristic layer. The second heuristic layer subsumes (overrides decisions made in) the first, and forms the most significant part of system and makes use of semantic knowledge in WordNet with regards to the head of the PP's component noun phrase in order to guide semantic classification. Specifically, depending on the $NP$ expansion rule that expands the direct object of the preposition, the relevant head of the noun phrase is extracted. For instance, for pronouns and proper $NPs$, the whole $NP$ is considered, otherwise, only the grammatic head of the noun phrase (which has been annotated automatically) is considered. Given the relevant head of the noun phrase, hypernyms for that word is looked up in WordNet using the command `\texttt{\footnotesize wn \$noun -hypen}'. This will often result in multiple possible hypernym expansions due to the slightest polysemy of the noun phrase, so only the first few senses% \footnote{In WordNet, this translates as being the few most frequent senses, as determined through corpus-based training.% } up to some threshold (4 or 5) are considered. The hypernym trees are then searched in decreasing frequency of sense for particular keywords indicating semantic role, for instance the hypernym {}``\texttt{\footnotesize time period}'', which can be derived from the noun {}``\texttt{\footnotesize yesterday}'', indicates a temporal relation. This search is done in such a way that the most common semantic categories are given priority. This approach is similar to (and inspired by!) approaches using FrameNet as the semantic resource, and an evaluation of this technique follows in Section \ref{sub:Evaluation-of-Automatic}. \subsection{\label{sub:PP-Attachment}PP Attachment} Although prepositional phrases {}``... can appear within all the other major phrase types'' \cite{manning2002}, the task of achieving automatic PP-attachment syntactically has traditionally required attachment either to a verb phrase or to a noun phrase, historically accomplished by means of a specific binary decision between two particular methods: \begin{lyxlist}{00.00.0000} \item [{Right~Association:}] A constituent tends to attach to another immediately to its right. This approach favours attachment to the noun \cite{kimball73}. \\ (ex. \texttt{\scriptsize {}``I (VP saw (NP the dog (PP with its puppies ))''} ) \item [{Minimal~Attachment:}] A constituent tends to attach to existing nonterminals using the fewest intermediate nodes. This approach favours attachment to the verb \cite{frzier78}.\\ (ex. \texttt{\scriptsize {}``I (VP saw (NP the dog) (PP with my binoculars))''} ) \end{lyxlist} However, Whittemore et al. \cite{whittemore89}, showed that \emph{neither} of these complementary tactics can account for more than 55\% of cases in general texts - and that each are poor predictors of how people resolve ambiguity \cite{mitchell2002}. The first solution to this shortcoming is to involve statistics and lexical knowledge. It has been shown in Brill and Resnik (1994) \cite{brill94}.{\footnotesize{} }{}``\texttt{\footnotesize I saw the fool on the hill with the telescope}'', where there exist numerous examples of ambiguity, but our concern is with the ambiguity of PP-attachment. Does the prepositional phrase {}``\texttt{\footnotesize on the hill}'' refer to the location of the fool, or of the speaker? Does the prepositional phrase {}``\texttt{\footnotesize with the telescop}e'' attach to the verb \texttt{\footnotesize {}``saw}'', or to either of the nouns {}``\texttt{\footnotesize the fool}'' or {}``\texttt{\footnotesize the hill}''? Both of these suggested improvements to a syntactic approach to PP-attachment - corpus-based statistical learning and world knowledge (with inference), are beyond the scope of this project. \subsection{\label{sub:Consideration-of-Semantics}Consideration of Semantics} Determening semantics (Level 4) is beyond the scope of this work presentely due to time constraints. However, we present a few general ideas on how that could possibly be done. One (rather shallow) way is to build subcategories of the 7 classes we have used. For example, locative-type PPs may have lexical entries with semantic relations, such as ``part-to-whole'', ``betweenness'', and ``relative distance'' relations, which build up ``path'' and ``orientation'' structures for ``movement-directional'' and ``stative-locational'' interpretations according to \cite{locativesem}. Similarly to VPs, we'll have to look at the (semantics of) argument structure of such PPs. In general, a hierarchy of subcategories of the original classes of PPs will have to be implemented. That would possibly require more than two passes of over the parses we currently have. Tests for argument structure can be done as presented in \cite{verspoor-perspective}. \section{Experiments \& Analysis} In order to gauge the effectiveness of our heuristics-based approach, the evaluation of our system on corpora, and the subsequent experimental analysis must be performed. Such experiment involves a `training' phase where the syntactic and semantic characteristics of the prepositional phrases are analyzed and represented by heuristics, and a `testing' phase where the efficacy of those heuristics are measured statistically. The process is described in the following subsections. \subsection{Experimental Setup} \subsubsection{Corpora} For the purposes of experiment, a collection of texts from the Wall Street Journal (WSJ) corpus are chosen at random to be representative of the available corpora, and are divided into two sets: The first is comprised of \textbf{20} texts (\emph{\textasciitilde{}74\%}) used for `\textbf{training}' of the system - that is, our heuristics are modified to best suit these documents. The second set is comprised of \textbf{7} texts (\emph{\textasciitilde{}26\%}) used for `\textbf{testing}' the system - from which our empirical measures of performance are derived. The breakdown is shown in Tables \ref{cap:Training-corpora} and \ref{cap:Test-corpora}. \begin{table} \begin{centering} \begin{tabular}{|c|c|c|c|} \hline \texttt{\tiny 891027-0018.txt} & \texttt{\tiny 891027-0023.txt} & \texttt{\tiny 891027-0038.txt} & \texttt{\tiny 891027-0047.txt}\tabularnewline \hline \texttt{\tiny 891027-0056.txt} & \texttt{\tiny 891027-0066.txt} & \texttt{\tiny 891027-0082.txt} & \texttt{\tiny 891027-0090.txt}\tabularnewline \hline \texttt{\tiny 891027-0101.txt} & \texttt{\tiny 891027-0103.txt} & \texttt{\tiny 891027-0114.txt} & \texttt{\tiny 891027-0172.txt}\tabularnewline \hline \texttt{\tiny 891027-0184.txt} & \texttt{\tiny 891030-0008.txt} & \texttt{\tiny 891030-0011.txt} & \texttt{\tiny 891030-0020.txt}\tabularnewline \hline \texttt{\tiny 891030-0028.txt} & \texttt{\tiny 891030-0037.txt} & \texttt{\tiny 891030-0066.txt} & \texttt{\tiny 891030-0085.txt}\tabularnewline \hline \end{tabular} \par\end{centering} \caption{\label{cap:Training-corpora}Training documents from WSJ corpus} \end{table} \begin{table} \begin{centering} \begin{tabular}{|c|c|c|c|} \hline \texttt{\tiny 891027-0007.txt} & \texttt{\tiny 891027-0040.txt} & \texttt{\tiny 891027-0081.txt} & \texttt{\tiny 891027-0099.txt}\tabularnewline \hline \texttt{\tiny 891027-0111.txt} & \texttt{\tiny 891030-0019.txt} & \texttt{\tiny 891030-0093.txt} & \texttt{\tiny -}\tabularnewline \hline \end{tabular} \par\end{centering} \caption{\label{cap:Test-corpora}Testing documents from WSJ corpus} \end{table} \subsubsection{Human Annotation} All 27 documents were hand-annotated for the occurence and grammatical structure of prepositional phrases. This task was facilitated by two factors: \begin{itemize} \item Only prepositional phrases (and their constituents) were annotated, saving the annotators the trouble of doing full sentence parsing. \item As a consequence of the first point, PP-attachment was not taken into consideration. \end{itemize} Subsequently the parses were augmented with the semantic role annotations from Section \ref{sub:Semantic-Role-Annotations}. In cases where the prepositional phrase did \emph{not} fall into one of the 7 semantic categories, the tag was left simply as \textbf{PP}, without additional markup. The annotation of the 20 training documents guided the production of rules for grammatical expansion and heuristics for semantic role markup. The annotation of the 7 test documents allowed for numeric performance measures, and statistical evaluation. \subsubsection{Measures of Performance} Two principal measures of performance are \textbf{precision} and \textbf{recall}, defined thusly \cite{quirk85}: \begin{lyxlist}{00.00.0000} \item [{Precision:}] $P=\frac{\sharp\, of\, correct\, answers\, given\, by\, the\, system}{\sharp\, of\, answers\, given\, by\, the\, system}$ \item [{Recall:}] $R=\frac{\sharp\, of\, correct\, answers\, given\, by\, the\, system}{total\,\sharp\, of\, possible\, answers\, in\, the\, text}$ \end{lyxlist} For each of these measures, an `answer' can be defined with regards to the task at hand, and will be specified in the Section \ref{sub:Results-&-Analysis}. For instance, when measuring the bare PP-chunker grammar, recall would be defined as the total number of correctly identified prepositional phrases among all actual prepositional phrases in a text. An additional statistical measure is the F-Measure, which describes the combined measure of precision and recall, and is described by \[ F=\frac{\left(\beta^{2}+1\right)PR}{\beta^{2}P+R}\] where we consider $\beta=1$ (equal weight to precision and recall). Non-statistical measures of performance include computational measures and the qualitative evaluation of rules. \subsection{\label{sub:Results-&-Analysis}Results \& Analysis} \subsubsection{Evaluation of PP-grammar} The discovery of minimum non-characterized PPs, which comprises the first level of the system, is first analyzed independently, since its performance will drastically affect the performance of later stages. The first pass of evaluation leads to a recall of $R=\frac{302}{369}\approx81.8$\% and a precision of $P=\frac{302}{409}\approx73.8$\%. The relatively poor precision is quickly discovered to be a result of poor preprocessing of the WSJ texts, which results in errors in tagging. Specifically, erroneous prepositional phrases are found in headers of the WSJ texts and at the ends of sentences at at points where punctuation causes difficulty to Brill's Tagger, where errors are likely made during the transformational tagging stage. With additional preprocessing steps for those circumstances listed, recall improved to $R=\frac{302}{369}\approx81.8$\% and precision to $P=\frac{302}{381}\approx79.3$\%. Though a significant improvement, additional improvements are possible. Analysing the training corpus, we find that many potential prepositional phrases \emph{will not} parse due to poor parsing of more complex noun phrases which would normally form the object of the PP. Furthermore, the rule $PP\Leftarrow IN\, RB$, which was originally included to deal with phrases of the form {}``\texttt{\scriptsize in particular}'', was overgenerating many parses. Modifying the rules for $NPs$ to take into consideration aggragate NPs, and removing the rule $PP\Leftarrow IN\, RB$ results in a final recall of $R=\frac{324}{369}\approx87.8$\% and precision to $P=\frac{324}{381}\approx85.0$\%, which translates into an F-Measure of $F\approx86.4$\%. Evidently, relatively high precision and recall can be achieved for prepositional phrases with relatively few grammatical rules. This is in stark contrast to the automatic noun phrase or verb phrase chunkers,which traditionally score lower and require more numerous and more complexrules. This tends to validate the theory in the literature which paints PPs ashaving a basic syntactic structure, as described in Section \ref{sub:Discovery-of-Minimal}. The improvements made at each evaluation pass are summarized in Figure \ref{cap:Precision-and-recall-PP-chunker}. \begin{figure} \begin{centering} \includegraphics[width=\columnwidth]{PP-eval.pdf} \par\end{centering} \caption{\label{cap:Precision-and-recall-PP-chunker}Precision and recall for evaluation of PP-chunker.} \end{figure} \subsubsection{\label{sub:Evaluation-of-Automatic}Evaluation of Automatic Categorization} Utilizing lexical knowledge first, then subsuming it with semantic knowledge,leads to very encouraging results. Measuring the recall, precision, and $F$-measure as previously defined for all 7 test, we break down the analysis into the 4 most frequent catgories, where un-annotated PPs are considered towards the total positively if they do not strictly fit into our seven categories, and negatively otherwise% \footnote{So total fractions do not necessarily equal the sum of its parts.% }. This breakdown is shown in Table \ref{cap:Evaluation-of-Automatic}. \begin{table} \begin{centering} \begin{tabular}{|c||c|c|c|} \hline & {\tiny Recall} & {\tiny Precision} & {\tiny $F$-measure}\tabularnewline \hline {\tiny Locative PPs only} & {\tiny $\frac{128}{147}=87.1$\%} & {\tiny $\frac{128}{143}=89.5$\%} & {\tiny 88.3\%}\tabularnewline \hline {\tiny Temporal PPs only} & {\tiny $\frac{78}{98}=79.6$\%} & {\tiny $\frac{78}{96}=81.3$\%} & {\tiny 80.5\%}\tabularnewline \hline {\tiny Directional PPs only} & {\tiny $\frac{68}{79}=86.1$\%} & {\tiny $\frac{68}{74}=91.9$\%} & {\tiny 89.0\%}\tabularnewline \hline {\tiny Manner PPs only} & {\tiny $\frac{17}{26}=65.4$\%} & {\tiny $\frac{17}{28}=60.7$\%} & {\tiny 63.1\%}\tabularnewline \hline {\tiny Total} & {\tiny $\frac{293}{369}=79.4$\%} & {\tiny $\frac{293}{351}=83.5$\%} & {\tiny 81.5\%}\tabularnewline \hline \end{tabular} \par\end{centering} \caption{\label{cap:Evaluation-of-Automatic}Evaluation of Automatic Categorization} \end{table} It is immediately recognized that our heuristics seem exceptionally well-suited for directional-PPs, and exceptionally poorly suited to manner-PPs. Such a discrepancy can of course \emph{not} be linked merely to the heuristics for these two categories, since the system is not independent and heuristics for other categories play a significant role with each other. It should be noted that as more potential categories are taken under consideration, the more error we can expect due to the relative higher rate of perplexity. \section{Future Work} Most of the future work would be dedicated to the Levels 3 and 4 outlined at the beginning, i.e. PP-attachment and Semantic Analysis of PPs. Both problems are quite difficult to solve and will require a lot of research-trial-and-error attempts of the existing and new proposals. New, more robust, tools will be required, such as stochastic methods in NLP and a lot more training and testing data. For example, we could assign probabilities to our grammar rules and encode the attachment information into the syntactic productions. Likewise, the detection of semantic roles of PPs and their arguments can be done heuristically and statistically with two or more passes. \label{sect:bib}\bibliographystyle{alpha}
\section{Introduction} \label{sec:intro} 2MASS\,J05352184$-$0546085 (2M0535$-$05) is a benchmark object for brown dwarf (BD) science since it offers the rare opportunity of independent radius and mass measurements on substellar objects. The observed values constrain evolutionary and structural models \citep{1997MmSAI..68..807D, 1998A&A...337..403B, 2000ARA&A..38..337C, 2002A&A...382..563B, 2007A&A...472L..17C}. 2M0535$-$05 is located in the Orion Nebulae, a star-forming region with an age of 1 ($\pm 0.5$)\,Myr. If both components formed together, as commonly believed, then this system allows for effective temperature ($T_{\mathrm{eff}}$) and luminosity ($L$) measurements of two BDs at the same age. However, this system is observed to have an unexpected temperature reversal \citep{2006Natur.440..311S}, contravening theoretical simulations: the more massive component (the primary) is the cooler one. From the transit light curve, the ratio of the effective temperatures can be accurately determined to $T_{\mathrm{eff},2}/T_{\mathrm{eff},1} = 1.050 \pm 0.002$ \citep{2009ApJ...697..713M, 2009ApJ...699.1196G}. From spectroscopic measurements then, the absolute values can be constrained. The primary, predicted to have $T_{\mathrm{eff},1} \approx 2\,870$\,K \citep{1998A&A...337..403B}, has an observed value of $\approx 2\,700$\,K, whereas the surface temperature of the secondary, predicted to be $T_{\mathrm{eff},2} \approx 2\,750$\,K, is most compatible with $T_{\mathrm{eff},2} \approx 2\,890$\,K. One explanation for the temperature discrepancies is suppression of convection due to spots on the surface of the primary. If a portion of a BD's surface is covered by spots, its apparent temperature will be reduced, resulting in an increase in the estimated radius in order for the measured and expected luminosities to agree \citep{2007A&A...472L..17C}. With a spot coverage of 30 - 50\% and a mixing length parameter $\alpha = 1$ most of the mismatches between predicted and observed radii for low-mass stars (LMS) can be explained \citep{2008MmSAI..79..562R}. Observations of spots on both of the 2M0535$-$05 components \citep{2009ApJ...699.1196G}, as inferred from periodic variations in the light curve, and measurements on the H$_{\alpha}$ line of the combined spectrum during the radial velocity maxima \citep{2007ApJ...671L.149R} suggest that enhanced magnetic activity and the accompanying spots on the primary indeed play a key role for its temperature deviation. But even if the spot coverage on the primary serves as an explanation for the primary's reduced $T_{\mathrm{eff}}$, the secondary's luminosity overshoot of $\approx 2.3 \cdot 10^{24}$\,W, as compared to the \citet{1998A&A...337..403B} models, suggests some additional processes may be at work. The temperature reversal between the primary and secondary may result from a difference between their ages. The secondary could be $\approx 0.5$\,Myr older than the primary, as proposed by \citet{2007ApJ...664.1154S} (see also \citet{1997MmSAI..68..807D}). A difference of 0.5\,Myr could allow the secondary to have converted the necessary amount of gravitational energy into heat\footnote{In contrast to the \citet{1998A&A...337..403B} tracks, the models by \citet{1997MmSAI..68..807D} predict a temperature increase in BDs for the first $\approx 30$\,Myr of their existence.}, which would explain its luminosity excess. But evolutionary models are very uncertain for ages $\lesssim 1$\,Myr \citep{2002A&A...382..563B, 2005AN....326..905W, 2007ApJ...655..541M, 2007IAUS..239..197M} and, in any case, the age determination and physical natures of these very young objects is subject to debate \citep{2008Natur.453.1079S, 2009IAUS..258..161S}. Furthermore, the mutual capture of BDs and LMS into binary systems after each component formed independently is probably too infrequent to account for the large number of eclipsing LMS binaries with either temperature reversals or inflated radii \citep{2001A&A...366..965G, 2007JSARA...1....7C, 2008MmSAI..79..562R, 2009NewA...14..496C, 2009ApJ...691.1400M}. Here, we consider the role that tidal heating may play in determining the temperatures of the BDs. In Table \ref{tab:parameters} we show the parameters of 2M0535$-$05 necessary for our calculations. The computed energy rates will add to the luminosity of the BDs in some way (Sect. \ref{sub:converting}) and will contribute to a temperature deviation compared to the case without a perturbing body (Sect. \ref{sec:results}). All these energy rates must be seen in the context of the luminosities of the BDs: $L_1 \approx 8.9 \cdot 10^{24}$\,W (luminosity of the primary) and $L_2 \approx 6.6 \cdot 10^{24}$\,W (luminosity of the secondary). At a distance $a$ to the primary component, its luminosity is distributed onto a sphere with area $4\,\pi\,a^2$. The secondary has an effective -- i.e. a 2D-projected -- area of $\pi\,R_2^2$. With $F_{1,\mathrm{a}}$ as the flux of the primary at distance $a$, the irradiation from the primary onto the secondary $L_{1 \rightarrow 2}$ is thus given by \begin{equation}\label{equ:lum_irrad} L_{1 \rightarrow 2} = \pi \ R_2^2 \ F_{1,\mathrm{a}} = \pi \ R_2^2 \ \frac{L_1}{4 \ \pi \ a^2} = L_1 \frac{R_2^2}{4 \ a^2} . \end{equation} \noindent Using that equation, we calculate the mutual irradiation of the BDs: $L_{1 \rightarrow 2} \approx 8.5 \cdot 10^{21}$\,W and $L_{2 \rightarrow 1} \approx 1.0 \cdot 10^{22}$\,W. These energy rates are two and three orders of magnitude lower, respectively, than the observed luminosity discrepancy. Hence, we assume that mutual irradiation can be ignored. This simplification is in contrast to the cases of the potentially inflated transiting extrasolar planets WASP-4b, WASP-6b, WASP-12b, and TrES-4, where stellar irradiation \citep{2009arXiv0910.4394I} dominates tidal heating by several magnitudes. Various tidal models haven been used to calculate tidal heating in exoplanets \citep{2001ApJ...548..466B, 2008MNRAS.391..237J, 2008ApJ...681.1631J, 2009ApJ...695.1006B}, which may in fact be responsible for previous discrepancies between interior models and radii of transiting exoplanets \citep{2008MNRAS.391..237J, 2008ApJ...681.1631J, 2009ApJ...700.1921I}. This success in exoplanets motivates our investigation into BDs. While many different tidal models are available, there is no consensus as to which is the best. For this reason, we apply a potpourri of well-established models to the case of 2M0535$-$05 in order to compare the different results. As we show, tidal heating may account for the temperature reversal and it may have a profound effect on the longer-term thermal evolution of the system. The coincidence of $P_{\mathrm{orb}} / P_1 \approx 2.9698 \approx 3$, with $P_{\mathrm{orb}}$ as the orbital and $P_1$ primary's rotation period, has been noted before but we assume no resonance between the primary's rotation and the orbit for our calculations. These resonances typically occur in systems with rigid bodies where a fixed deformation of at least one body persists, such as in the Sun-Mercury configuration with Mercury trapped in a 3/2 spin-orbit resonance. We assume that, in the context of tides, BDs may rather be treated as fluids and the shape of the body is not fixed. With this paper, we present the first investigation of tidal interaction between BDs. In Sect. \ref{sec:tid_mod} we introduce four models for tidal interaction and discuss how we convert the computed energy rates into an increase in effective temperature. Sect. \ref{sec:results} is devoted to the results of our calculations, while we deal with the observational implications in Sect. \ref{sec:discussion}. We end with conclusions about tidal heating in 2M0535$-$05, and in BDs in general, in Sect. \ref{sec:conclusions}. \section{Tidal Models} \label{sec:tid_mod} Two qualitatively different models of tidal dissipation and evolution have been developed over the last century: The `constant-phase-lag' \citep[][Wis08 and FM08 in the following]{1966Icar....5..375G, 2008Icar..193..637W, 2008CeMDA.101..171F}, and the `constant-time-lag' model \citep[][Hut81 in the following]{1981A&A....99..126H}. In the former model, the forces acting on the deformed body are described by a superposition of a static equilibrium potential and a disturbing potential (FM08). The latter model assumes the time between the passage of the perturbing body overhead and the passage of the tidal bulge is constant. Although both models have been used extensively, it is not clear which model provides a more accurate description of the effects of tides, so we apply formulations of both models. In the `constant-phase-lag' model of FM08, quantitative expressions have been developed to second order in eccentricity $e$ while the others include also higher orders. Higher and higher order expansions require assumptions about the dependence of a body's tidal response to an increasing number of tidal frequencies, which involves considerable uncertainty. Therefore higher order expansions do not necessarily provide more accuracy \citep[FM08;][]{2009ApJ...698L..42G}. In the `constant-phase-lag' model of Wis08, expressions in $e$ are developed to $8^{\mathrm{th}}$ order. The `constant-time-lag' model of Hut81 does not include possible obliquities, while an enhanced version of that model by \citet{2007A&A...462L...5L} (Lev07) does. Tidal dissipation in BDs has not been observed or even considered previously, and hence, neither model should take precedence when calculating their tidal dissipation, especially since neither tidal model is definitive \citep{2009ApJ...698L..42G}. As our investigation is the first to consider tidal effects on BDs, we will employ several applicable, previously published models to 2M0535$-$05. By surveying a range of plausible models and internal properties, usually encapsulated in the `tidal dissipation function' $Q$ \citep{1966Icar....5..375G}, we may actually be able to determine which model is more applicable to the case of BDs -- assuming, of course, that tidal dissipation contributes crucially to the observed temperature inversion. \renewcommand{\arraystretch}{1.3} \begin{table}[h] \centering \caption{Orbital and physical parameters of 2M0535$-$05} \label{tab:parameters} \begin{tabular}{l|c} \hline \hline \hline \hline \multicolumn{1}{c}{\textsc{property} }& \textsc{observed value}\\ \hline $a$, semi-major axis$^1$ & 0.0407 $\pm 0.0008$\,AU \\ $e$, eccentricity$^1$ & 0.3216 $\pm 0.0019$ \\ $P_{\mathrm{orb}}$, orbital period$^1$ & 9.779556 $\pm 0.000019$\,d \\ $i$, orbital inclination to the line of sight$^1$ & 88.49 $\pm 0.06^\circ$ \\ age$^1$ & 1 $\pm 0.5$\,Myr \\ $T_{\mathrm{eff},1}$, primary effective temperature$^1$ & 2\,715 $\pm 100$\,K \\ $T_{\mathrm{eff},2}/T_{\mathrm{eff},1}$, effective temperature ratio$^1$ & 1.050 $\pm 0.002$ \\ $M_{1}$, primary mass$^1$ & 0.0572 $\pm 0.0033$\,$M_{\odot}$ \\ $M_{2}$, secondary mass$^1$ & 0.0366 $\pm 0.0022$\,$M_{\odot}$ \\ $R_{1}$, primary radius$^1$ & 0.690 $\pm 0.011$\,$R_{\odot}$ \\ $R_{2}$, secondary radius$^1$ & 0.540 $\pm 0.009$\,$R_{\odot}$ \\ $L_{1}$, primary luminosity$^3$ & $8.9 \cdot 10^{24}$ $\pm 3 \cdot 10^{24}$\,W \\ $L_{2}$, secondary luminosity$^3$ & $6.6 \cdot 10^{24}$ $\pm 2 \cdot 10^{24} $\,W \\ $P_{1}$, rotational period of the primary$^1$ & 3.293 $\pm 0.001$\,d \\ $P_{2}$, rotational period of the secondary$^1$ & 14.05 $\pm 0.05$\,d \\ $ \bar{T}_{\mathrm{eff},1}$, modeled $T_{\mathrm{eff}}$ for the primary$^2$ & 2\,850\,K\\ $ \bar{T}_{\mathrm{eff},2}$, modeled $T_{\mathrm{eff}}$ for the secondary$^2$ & 2\,700\,K\\ $ \bar{R}_1$, modeled radius for the primary$^2$ & $0.626 \, R_{\odot}$\\ $ \bar{R}_2$, modeled radius for the secondary$^2$ & $0.44 \, R_{\odot}$\\ \hline \multicolumn{2}{l}{}\\ \multicolumn{2}{l}{$^1$ \citet{2009ApJ...699.1196G}, $^2$ \citet{1998A&A...337..403B},}\\ \multicolumn{2}{l}{$^3$ assuming an uncertainty of 200\,K in $T_{\mathrm{eff},1}$ and $T_{\mathrm{eff},2}$} \end{tabular} \end{table} \subsection{Constant phase lag} \label{sub:const_ang_lag} \subsubsection{Tidal model \#1} \label{subsub:tid_model_1} The potential of the perturbed body can be treated as the superposition of periodic contributions of tidal frequencies at different phase lags and the expression for the potential can be expanded to first order in those lags (FM08). Those phase lags $\varepsilon_{k,i \ | \ k = 0, 1, 2, 5, 8, 9}$ of the $i^{\mathrm{th}}$ body that we will need for our equations are given by \begin{align}\label{equ:epsilon} \nonumber Q_i \ \varepsilon_{0,i} & = \Sigma(2 \Omega_i - 2 n)\\ \nonumber Q_i \ \varepsilon_{1,i} & = \Sigma(2 \Omega_i - 3 n)\\ \nonumber Q_i \ \varepsilon_{2,i} & = \Sigma(2 \Omega_i - n)\\ \nonumber Q_i \ \varepsilon_{5,i} & = \Sigma(n)\\ \nonumber Q_i \ \varepsilon_{8,i} & = \Sigma(\Omega_i - 2 n)\\ Q_i \ \varepsilon_{9,i} & = \Sigma(\Omega_i) \hspace{1.5cm} i \in \{1,2\} ,\\ \nonumber \end{align} \noindent where $\Sigma(x)$ is the algebraic sign of $x$, thus $\Sigma(x)~=~+1~\vee~-1$, $n = 2 \pi / P_{\mathrm{orb}}$ is the orbital frequency and $\Omega_i = 2 \pi / P_i$ are the rotational frequencies of the primary ($i = 1$) and secondary ($i = 2$), $P_i$ being their rotational periods. The tidal frequencies are functions of the tidal quality factor $Q$ of the deformed object, which parametrizes the object's tidal response to the perturber. It is defined as \begin{equation} Q^{-1} = \frac{1}{2 \pi E_0} \int_0^{P_{\mathrm{orb}}} \mathrm{d}t \left( - \frac{\mathrm{d}E}{\mathrm{d}t} \right) , \end{equation} \noindent where $E_0$ is the maximum energy stored in the tidal distortion and the integral over the energy dissipation rate $-$d$E$/d$t$ is the energy lost during one orbital cycle \citep{1966Icar....5..375G}. Although \citet{2004ApJ...610..477O} conclude that tidal dissipation rates of giant planets are not adequately represented by a constant $Q$-value, many parameterized tidal models rely on this quantity. Measurements of the heat flux from Jupiter's moon Io during the fly-by of the Voyager 1 spacecraft, combined with a specific model of the history of the orbital resonance, allowed for an estimate for the quality factor $Q_{\jupiter}$ of Jupiter to be $2 \cdot 10^{5} < Q_{\jupiter} < 2 \cdot 10^{6}$ \citep{1979Natur.279..767Y} while \citet{2001AJ....122.2734A} used historical changes in Io's orbit to infer that $Q_{\jupiter}$ is around $10^{5.3}$. However, \citet{2008DPS....40.0403G} pointed out that $Q = \infty$ is not ruled out \citep[see also][]{1980LPI....11..871P, 1993ApJ...406..266I}. Tides raised by Neptune on its moons help to constrain the planet's quality factor to $10^{3.95} < Q_{\neptune} < 10^{4.56}$ \citep{2008Icar..193..267Z}. For M dwarfs, $Q_{\mathrm{dM}}$ is assumed to be of order $10^5$, whereas for rigid bodies like Earth $20 \lesssim Q \lesssim 500$ \cite[][and references therein]{2001GeoJI.144..471R, 2004ApJ...614..955M}. For BDs, however, $Q$ is even more uncertain, thus we will handle it as a free parameter in our procedures. FM08 allows for the tidal amplitude to be different from what it would be if the tide-raising body were fixed in space. This concern is met by the dynamical Love number $k_\mathrm{d}$ under the assumption that the tidally disturbed body had infinite time to respond. Without better knowledge of a body's response to tides, we assume the dynamical Love number is the same as the potential Love number of degree 2, $k_2$. For the gas planets of the solar system, this number has been calculated by \citet{1977Icar...32..443G}. BDs may rather be treated as polytropes of order $\mathfrak{n} = 3/2$ (I. Baraffe, private communication). We infer the Love number from the relation $k_\mathrm{2} = 2 k_\mathrm{aps}$ \citep{2002ApJ...573..829M} and use the tables of apsidal motion constants $k_\mathrm{aps}$ given in \citet{1955MNRAS.115..101B}. These authors provide numerical calculations for $k_{\mathrm{aps}}$ for a polytrope of $\mathfrak{n} = 3/2$. We find $k_\mathrm{aps} = 0.143$ and thus $k_\mathrm{d} \equiv k_\mathrm{2} = 0.286$. This places $k_2$ for BDs well in the regime spanned by the gas giants of the solar system: Jupiter ($k_2 = 0.379$), Saturn ($k_2 = 0.341$), Uranus ($k_2 = 0.104$) and Neptune ($k_2 = 0.127$) \citep{1977Icar...32..443G}. Before we proceed to the equations for the tidal heating rates, we sum up those for the orbital evolution of the semi-major axis $a$, the eccentricity $e$ and the putative obliquity $\psi$. The latter parameter is the angle between the equatorial plane of one of the two bodies in a binary system and the orbital plane \citep{2005ApJ...631.1215W}, frequently referred to as spin-orbit misalignment. We use Eqs. (56), (60) and (61) from FM08 but our equations for a binary system with comparable masses need slight modifications since both constituents contribute significantly to the evolution of $a$ and $e$. We add both the terms for the secondary being the perturber of the primary ($i = 1$, $j = 2$) and vice versa, since only spin-orbit coupling is relevant, whereas spin-spin interaction can be neglected. This results in \begin{align}\label{equ:a_FM} \nonumber \frac{\mathrm{d}a}{\mathrm{d}t} = \sum_{ \substack{i \, = \, 1,2 \\ i \, \neq \, j} } \frac{3 k_{\mathrm{d},i} M_j R_i^5 n}{4 M_i a^4} \ ( \ &4 \varepsilon_{0,i} + e^2 [ - 20 \varepsilon_{0,i} + \frac{147}{2} \varepsilon_{1,i} + \frac{1}{2} \varepsilon_{2,i}\\ &- 3 \varepsilon_{5,i}] - 4 S_i^2 [\varepsilon_{0,i} - \varepsilon_{8,i}] \ ) , \end{align} \begin{equation}\label{equ:e_FM} \frac{\mathrm{d}e}{\mathrm{d}t} = - \sum_{ \substack{i \, = \, 1,2 \\ i \, \neq \, j} } \frac{3 e k_{\mathrm{d},i} M_j R_i^5 n}{8 M_i a^5} \left( 2 \varepsilon_{0,i} - \frac{49}{2} \varepsilon_{1,i} + \frac{1}{2} \varepsilon_{2,i} + 3 \varepsilon_{5,i} \right) , \end{equation} \begin{equation}\label{equ:psi_FM} \frac{\mathrm{d}\psi_i}{\mathrm{d}t} = \frac{3 k_{\mathrm{d},i} M_j R_i^5 n}{4 M_i a^5} \ S_i \ \left( - \varepsilon_{0,i} + \varepsilon_{8,i} + - \varepsilon_{9,i} \right) , \end{equation} \noindent where $k_{\mathrm{d},i}$ is the dynamical Love number, $M_i$ the mass and $R_i$ the radius of the deformed BD, $S_i \coloneqq \sin(\psi_i)$, with $\psi_i$ as the obliquity of the perturbed body, and $\varepsilon_{k,i \ | \ k = 0, 1, 2, 5, 8, 9}$ are the tidal phase lags, given in Eq. (\ref{equ:epsilon}). The total energy that is dissipated within the perturbed body, its tidal energy rate, can be determined by summing the work done by tidal torques (Eqs. (48) and (49) in FM08). The change in orbital energy of the $i^{\mathrm{th}}$ body due to the $j^{\mathrm{th}}$ body is given by \begin{align}\label{equ:E_orb_mod1} \nonumber \dot{E}_{\mathrm{orb},i}^{\mathrm{\#1}} = \ \underbrace{ \frac{3 k_{\mathrm{d},i} G M_j^2 R_i^5}{8 a^6} }_p \ n \ ( \ &4 \varepsilon_{0,i} + e^2 [-20 \varepsilon_{0,i} + \frac{147}{2} \varepsilon_{1,i} + \frac{1}{2} \varepsilon_{2,i}\\ &- 3 \varepsilon_{5,i} ] - 4 S_i^2 \ [\varepsilon_{0,i} - \varepsilon_{8,i}] \ ) \end{align} \noindent and the change in rotational energy is deduced to be \begin{align}\label{equ:E_rot_mod1} \nonumber \dot{E}_{\mathrm{rot},i}^{\mathrm{\#1}} = \ - \frac{3 k_{\mathrm{d},i} G M_j^2 R_i^5}{8 a^6} \ \Omega_i \ (& \ 4 \varepsilon_{0,i} + e^2 [-20 \varepsilon_{0,i} + 49 \varepsilon_{1,i} + \varepsilon_{2,i}]\\ & + 2 S_i^2 \ [- 2 \varepsilon_{0,i} + \varepsilon_{8,i} + \varepsilon_{9,i}] \ ) , \end{align} \noindent where $G$ is Newton's gravitational constant. The total energy released inside the body then is \begin{equation}\label{equ:E_tid_mod1} \dot{E}_{\mathrm{tid},i}^{\mathrm{\#1}} = - \ (\dot{E}_{\mathrm{orb},i}^{\mathrm{\#1}} + \dot{E}_{\mathrm{rot},i}^{\mathrm{\#1}}) > 0 . \end{equation} \noindent The greater-than sign in this equation is true, since either $\Omega_i < n$ and orbital energy is converted into rotational energy, or $\Omega_i > n$ and the body is decelerated by a transfer of rotational energy into orbital energy. In both cases, the dynamical energy of the system is released within the distorted body. For $\Omega_i = 0$, e.g., Eqs. (\ref{equ:E_orb_mod1}) and (\ref{equ:E_rot_mod1}) yield $\dot{E}_{\mathrm{orb},i}^{\mathrm{\#1}} = -p \cdot (4+57e^2+4S_i^2) / Q_i$ and $\dot{E}_{\mathrm{rot},i}^{\mathrm{\#1}} = 0$. The approach for the calculation of tidal energy rates with tidal model \#1 depends on processes due to non-synchronous rotation via $\varepsilon_{k,i} = \varepsilon_{k,i}(\Omega_i,n)$ and includes a putative obliquity $\psi_i$ and terms of $e$ up to the second order. After inserting the orbital and rotational periods for 2M0535$-$05, these equations reduce to \begin{align}\label{equ:E_tid_mod1_2M} \nonumber \dot{E}_{\mathrm{tid,1}}^{\mathrm{\#1}} & = \frac{3 k_{\mathrm{d},1} G M_2^2 R_1^5 }{8 Q_1 a^6} \ \left( \ [ 4 + 30 e^2 ] \Omega_1 - [4 + 51 e^2] n \ \right) ,\\ \dot{E}_{\mathrm{tid,2}}^{\mathrm{\#1}} & = \frac{3 k_{\mathrm{d},2} G M_1^2 R_2^5 }{8 Q_2 a^6} \ \left( \ [ 4 + 56 e^2 ] n + [2 S_2^2 - 4 - 28 e^2] \Omega_2 \ \right) . \end{align} \noindent Interestingly, for these particular values of $\Omega_1$, $\Omega_2$ and $n$, the $S_1$-terms for $\dot{E}_{\mathrm{tid},1}^{\mathrm{\#1}}$ cancel each other, so that it is not a function of $\psi_1$, whereas $\dot{E}_{\mathrm{tid},2}^{\mathrm{\#1}}$ does depend on $\psi_2$. \subsubsection{Tidal model \#2} The model of Wis08 includes terms in eccentricity up to the $8^\mathrm{th}$ order, predicting higher tidal energy rates than for the equations of model \#1. Equations for the evolution of the orbital parameters are not given in Wis08. Furthermore, in his theory the perturbed body is assumed to be synchronously rotating with the orbital period. Since this is not the case for either of the BDs in 2M0535$-$05, the following equations will only yield lower limits for the tidal heating. The tidal heating rates are given by \begin{equation}\label{equ:E_tid_mod2} \dot{E}_{\mathrm{tid},i}^{\mathrm{\#2}} =\frac{21 k_{2,i} G M_j^2 R_i^5 n}{2 Q_i a^6} \zeta_{\mathrm{Wis}}(e,\psi_i) \end{equation} \noindent with \begin{align}\label{equ:zeta_Wis} \nonumber \zeta_{\mathrm{Wis}}(e,\psi_i) = &\frac{2}{7} \frac{f_1^{\mathrm{Hut}}}{\beta^{15}} - \frac{4}{7} \frac{f_2^{\mathrm{Hut}}}{\beta^{12}} \ C_i + \frac{1}{7} \frac{f_5^{\mathrm{Hut}}}{\beta^9} \left(1+C_i^2\right)\\ & + \frac{3}{14} \frac{e^2 f_3^{\mathrm{Wis}}}{\beta^{13}} \ S_i^2 \cos(2 \Lambda_i) , \end{align} \noindent where we used $C_i \coloneqq \cos(\psi_i)$ and \begin{align}\label{equ:f_Hut_1} \nonumber \beta & = \sqrt{1-e^2} ,\\ \nonumber f_1^{\mathrm{Hut}} & = 1 + \frac{31}{2} e^2 + \frac{255}{8} e^4 + \frac{185}{16} e^6 + \frac{25}{64} e^8 ,\\ \nonumber f_2^{\mathrm{Hut}} & = 1 + \frac{15}{2} e^2 + \frac{45}{8} e^4 + \frac{5}{16} e^6 ,\\ \nonumber f_5^{\mathrm{Hut}} & = 1 + 3 e^2 + \frac{3}{8} e^4 ,\\ f_3^{\mathrm{Wis}} & = 1 - \frac{11}{6} e^2 + \frac{2}{3} e^4 + \frac{1}{6} e^6 ,\\ \nonumber \end{align} \noindent following the nomenclature of \citet{1981A&A....99..126H} and \citet{2008Icar..193..637W} as indicated. Furthermore, $k_{2,i}$ is the potential Love number of degree 2 for the $i^{\mathrm{th}}$ component of the binary system and $\Lambda_i$ is a measure of the longitude of the node of the body's equator on the orbit plane with respect to the pericenter of its orbit. In order to estimate the impact of $\Lambda_i$ in the last term in Eq. (\ref{equ:zeta_Wis}), we assume this impact to be as large as possible, $\Lambda_i = 0$, and compare it to the preceding terms. We find that for the case of 2M0535$-$05 the first three terms are of order 1, whereas the term connected to $\Lambda_i$ varies between $10^{-2}$ and $10^{-5}$, depending on $\psi_i$. These irrelevant contributions give us a justification to neglect the unknown values of $\Lambda_i$ in 2M0535$-$05 for our computations, facilitating the comparisons to the other models. \subsection{Constant time lag} \label{sub:const_time_lag} \subsubsection{Tidal model \#3} \label{subsub:tid_model_3} Instead of assuming phase lags and superposition of frequency-dependent potentials, the `equilibrium tide' model by \citet{1981A&A....99..126H} invokes a constant time lag $\tau$ between the line joining the centers of the two bodies and the culmination of the tidal bulge on the distorted object. With that assumption, the model of Hut81 is mutually exclusive with the assumption of a fixed angle lag \citep{1966Icar....5..375G}: in general, a fixed time lag and a fixed angle lag result in very different behaviors of the tidal bulge\footnote{If $e = 0$ and $\psi = 0$, then there is a single tidal lag angle $\varepsilon$ and the tidal dissipation funtion can be written as $Q = 1/\varepsilon = 1/(\tau n)$. For the course of an orbit, where the tidal evolution of $n$ is negligible, both $Q$ and $\tau$ can be fixed. However, in a general case where $\tau$ is constant in time, $Q$ will decrease as the orbital semi-major axis decays and $n$ increases. So $Q$ would not be constant.}. As for the case of the `constant-time-lag' model, we first sum up the equations governing the behavior of the orbital evolution. With the purpose of easing a comparison between Hut81's equations (Eqs. (9), (10), and (11) therein) and Eqs. (\ref{equ:a_FM}) - (\ref{equ:psi_FM}) from this paper for the theory of the `constant-phase-lag' model \#1, we transform the former into \begin{equation}\label{equ:a_Hut} \frac{\mathrm{d}a}{\mathrm{d}t} = \sum_{ \substack{i \, = \, 1,2 \\ i \, \neq \, j} } \frac{- 6 k_{\mathrm{aps},i} G M_j R_i^5}{a^7} \ \tau_i \ \left( 1 + \frac{M_j}{M_i} \right) \left( \frac{f_1^{\mathrm{Hut}}}{\beta^{15}} - \frac{f_2^{\mathrm{Hut}}}{\beta^{12}} \frac{\Omega_i}{n} \right) , \end{equation} \begin{equation}\label{equ:e_Hut} \frac{\mathrm{d}e}{\mathrm{d}t} = \sum_{ \substack{i \, = \, 1,2 \\ i \, \neq \, j} } \frac{- 27 k_{\mathrm{aps},i} G M_j R_i^5 e}{a^8} \ \tau_i \ \left( 1 + \frac{M_j}{M_i} \right) \left( \frac{f_3^{\mathrm{Hut}}}{\beta^{13}} - \frac{11}{18}\frac{f_4^{\mathrm{Hut}}}{\beta^{10}} \frac{\Omega_i}{n} \right) , \end{equation} \begin{align}\label{equ:psi_Hut} \nonumber \frac{\mathrm{d}\psi_i}{\mathrm{d}t} = &\frac{- 3 k_{\mathrm{aps},i} G M_j^2 R_i^3 \psi_i}{M_i a^6 r_{\mathrm{g},i}^2} \ \tau_i\\ & \times \left( \frac{f_2^{\mathrm{Hut}}}{\beta^{12}} \frac{n}{\Omega_i} - \frac{f_5^{\mathrm{Hut}}}{2 \beta^9} \left[ 1- \frac{r_{\mathrm{g},i}^2}{\beta} \frac{M_i + M_j}{M_j} \left( \frac{R_i}{a} \right)^2 \frac{\Omega_i}{n} \right] \right) , \end{align} \noindent with $k_{\mathrm{aps},i}$ as the apsidal motion constant of the perturbed body (see Sect. \ref{subsub:tid_model_1}), $r_{\mathrm{g},i}^2$ as the radius of gyration of the $i^{\mathrm{th}}$ body, which is defined by the body's moment of inertia $I_i = M_i r_{\mathrm{g},i}^2 R_i^2$, and \begin{align}\label{equ:f_Hut_2} \nonumber f_3^{\mathrm{Hut}} & = 1 + \frac{15}{4} e^2 + \frac{15}{8} e^4 + \frac{5}{64} e^6 ,\\ f_4^{\mathrm{Hut}} & = 1 + \frac{3}{2} e^2 + \frac{1}{8} e^4 . \end{align} \noindent Hut81 then calculates the energy dissipation rate within a binary system, caused by the influence of one of the two bodies on the other, as the change in the total energy $E = E_\mathrm{orb} + E_\mathrm{rot}$. Here, $E_\mathrm{orb}$ and $E_\mathrm{rot}$ are the orbital and rotational energies of the body (Eqs. (A28) - (A35) in Hut81). For the tidal heating rates of the $i^{\mathrm{th}}$ constituent within the binary, this yields \begin{equation}\label{equ:E_tid_mod3} \dot{E}_{\mathrm{tid},i}^{\mathrm{\#3}} = \frac{3 k_{\mathrm{aps},i} G M_j^2 R_i^5 n^2}{a^6} \ \tau_i \ \zeta_{\mathrm{Hut}}(e,\Omega_i,n) , \end{equation} \noindent where \begin{equation}\label{equ:zeta_Hut} \zeta_{\mathrm{Hut}}(e,\Omega_i,n) = \frac{f_1^{\mathrm{Hut}}}{\beta^{15}} - 2 \frac{f_2^{\mathrm{Hut}}}{\beta^{12}} \frac{\Omega_i}{n} + \frac{f_5^{\mathrm{Hut}}}{\beta^{9}} \frac{\Omega_i^2}{n^2} . \end{equation} \noindent Unfortunately, with these equations for the tidal energy rates model \#3 neglects a potential obliquity of the body, which prevents us from a direct comparison with the other tidal models. \subsubsection{Tidal model \#4} \label{subsub:tid_Lev} Lev07 extended Hut81's formula for the tidal energy rate to the case of an object in equilibrium rotation\footnote{Wis08 calls this `asymptotic nonsynchronous rotation'.} and they included possible obliquities \citep[see also][]{1997A&A...318..975N}, though they do not give the equations for the orbital evolution. Lev07's equations are equivalent to \begin{equation}\label{equ:E_tid_mod4} \dot{E}_{\mathrm{tid},i}^{\mathrm{\#4}} = \frac{3 k_{2,i} G M_j^2 R_i^5 n}{Q_{\mathrm{n},i} a^6} \ \zeta_{\mathrm{Lev}}(e,\psi_i) , \end{equation} \noindent where \begin{equation} \zeta_{\mathrm{Lev}}(e,\psi_i) = \frac{f_1^{\mathrm{Hut}}}{\beta^{15}} - \frac{ ( \ f_2^{\mathrm{Hut}} \ / \ \beta^{12} \ )^{2}}{f_5^{\mathrm{Hut}} \ / \ \beta^{9}} \ \left(1 + \frac{1}{1 - 2/S_i^2}\right) \end{equation} \noindent The `annual tidal quality factor' is given as $Q_{\mathrm{n}}^{-1} = n \ \tau$. Even though Lev07's equations invoke $Q_{\mathrm{n}}$ and their equations resemble those of the models with constant phase lag, their approach still assumes a constant-time-lag. Since Lev07 do not explicitly connect their $Q_{\mathrm{n}}$ to the $Q$ of FM08 (model \#1) and Wis08 (model \#2), we keep $Q$ and $Q_{\mathrm{n}}$ as two different constants for our further treatment. With these expansions, Eq. (\ref{equ:E_tid_mod4}) involves terms in eccentricity up to order $e^8$. But since model \#4 assumes tidal locking, i.e. $\dot{E}_{\mathrm{tid}}^{\mathrm{\#4}}$ is not a function of $\Omega$, this model also yields just a lower limit for the heating rates \citep{2008Icar..193..637W}. \subsection{Converting tidal heating into temperature increase} \label{sub:converting} Now that we have set up four distinct models for the calculations of the additional tidal heating term for the BDs, there are two physical processes that will be driven by these energy rates: tidal inflation and temperature increase. Let's take $\bar{L}$ as the luminosity of either of the two 2M0535$-$05 BDs that it would have if it were a single BD and $\bar{R}$ and $ \bar{T}_{\mathrm{eff}}$ as its corresponding radius and effective temperature. Then, by the Stefan-Boltzmann law \citep{Stefan, 1884AnP...258..291B} \begin{equation}\label{equ:StefanBotzlmann} \bar{L} = 4 \pi \bar{R}^2 \sigma_{\mathrm{SB}} \bar{T}_{\mathrm{eff}}^4 , \end{equation} \noindent where $ \sigma_{\mathrm{SB}}$ is the Stefan-Boltzmann constant. The radial expansion in the binary case is given by $\mathrm{d}R = R - \bar{R}$ and the temperature increase by $\mathrm{d}T = T_{\mathrm{eff}} - \bar{T}_{\mathrm{eff}}$. In its present state, the BD has a luminosity \begin{equation}\label{equ:L} L = \dot{E}_{\mathrm{in}} + \bar{L} , \end{equation} \noindent where $\dot{E}_{\mathrm{in}}$ is some additional internal energy rate. Solving Eq. (\ref{equ:L}) for the temperature increase yields: \begin{equation}\label{equ:dT1} \mathrm{d}T = \left( \frac{\dot{E}_{\mathrm{in}}}{4 \pi R^2 \sigma_{\mathrm{SB}}} + \left[ \frac{\bar{R}}{R} \right]^2 \bar{T}_{\mathrm{eff}}^4 \right)^{1/4} - \bar{T}_{\mathrm{eff}} . \end{equation} In the next step, we quantify the amount of tidal energy that is converted into internal energy, leading to an increase in effective temperature. Since we will use the virial theorem for an ideal, monoatomic gas to estimate the partition between internal and gravitational energy, we first have to assess the adequacy of treating the 2M0535$-$05 BDs as ideal gases. We therefore show the degeneracy parameter $\tilde{\Psi} = k_{\mathrm{B}} T / (k_{\mathrm{B}} T_{\mathrm{F}})$ as a function of radius in Fig.~\ref{fig:psi} \citep[][I. Baraffe, private communication.]{2000ARA&A..38..337C}. Here, $k_{\mathrm{B}}$ is the Boltzmann constant, $T$ is the local temperature within the gas and $E_{\mathrm{F}} = k_{\mathrm{B}} T_{\mathrm{F}}$ is the Fermi energy of a partially degenerate electron gas with an electron Fermi temperature $T_{\mathrm{F}}$. With respect to $M, T_\mathrm{eff}$ and $\log(g)$, $g$ being the body's gravitational acceleration at the surface, the BD structure model corresponds to that of the primary, but with an age of 4.9\,Myr. We find that for most of the BD, i.e. that portion of the structure in which the majority of the luminosity is released, $\tilde{\Psi}$ is of order 1. This means that we may indeed approximate the BDs as ideal gases. With the time derivative of the virial theorem for an ideal monoatomic gas \citep[][Sect. 3.1 therein]{1990sse..book.....K}, \begin{equation} L = \dot{E}_{\mathrm{in}} = - \dot{E}_{\mathrm{G}} / 2 , \end{equation} \noindent where $\dot{E}_{\mathrm{G}}$ is the temporal change in gravitational energy, we find that half of the additional tidal energy is converted into internal energy and the other half causes an expansion of the BD. There are currently no models for tidal inflation in BDs and the treatment is beyond the scope of this paper. Instead of including the modeled BD radii $\bar{R_i}$ into Eq. (\ref{equ:dT1}) we avoid further uncertainties and fix $\bar{R}/R = 1$ (see Sect. \ref{sec:conclusions} for a discussion of tidal inflation in the evolutionary context). The increase in effective temperature due to tidal heating then becomes \begin{equation}\label{equ:dT2} \mathrm{d}T = \left( \frac{\dot{E}_{\mathrm{tid}} / 2}{4 \pi R^2 \sigma_{\mathrm{SB}}} + \bar{T}_{\mathrm{eff}}^4 \right)^{1/4} - \bar{T}_{\mathrm{eff}} . \end{equation} \noindent For $\bar{T}_{\mathrm{eff},i}$ we took the values predicted by the \citet{1998A&A...337..403B} models (see Table \ref{tab:parameters}). Our neglect of tidal inflation makes this temperature increase an upper limit. Given that this neglect is arbitrary, we estimate how our constraints for $\log(Q_2) = 3.5$ and $\psi_2 = 50^\circ$ would change, if tidal inflation played a role in 2M0535$-$05. Comparing the observed radii of both BDs with the model predictions (see Table \ref{tab:parameters}), radial expansions of 10\% for the primary and 20\% for the secondary seem realistic. Theoretical investigations of tidal heating on the inflated transiting planet HD209458b by \citep{2009ApJ...700.1921I} support an estimate of tidal inflation by 20\%. As a test, we assumed that the secondary BD in 2M0535$-$05 is tidally inflated, where its radius in an isolated scenario would be 80\% of its current value, i.e. $\bar{R} = 0.8 \cdot R$ in Eq. (\ref{equ:dT1}). In the non-inflated scenario with $\bar{R}/R = 1$, the BD would reach a temperature increase of d$T = 60$\,K at $\log(Q_2) = 3.5$ and $\psi_2 = 50^\circ$ with model \#2 (see Sect. \ref{sub:tid_mod2_results}). With the inflation, however, $\log(Q_2) \approx 2.7$ is needed to achieve the same heating at $\psi_2 = 50^\circ$, whereas no obliquity at $\log(Q_2) = 3.5$ would yield significant heating. Thus, if tidal inflation in the secondary BD increases its radius by 20\%, then the value for the dissipation function required to yield the same $T_\mathrm{eff}$ would be about 0.8 smaller in $\log(Q)$ than in the case of no inflation. Therefore, the temperature we report in Sect. \ref{sec:results} may, at worst, correspond to $\log(Q)$ that is smaller by 0.8. \section{Results} \label{sec:results} \subsection{Orbital evolution} \label{sub:orb_evol} In order to get a rough impression of how far the orbital configuration of the system has evolved, we used the equations given in FM08, to compute the change of its eccentricity $e$ and of a possible obliquity $\psi_2$ of the secondary within the last 1.5\,Myr. Since this time span is the upper bound for the system's age, confined by its localization within the Orion Nebulae and indicated by comparison with BD evolutionary tracks, we thus get the strongest changes in $e$ and $\psi_2$. If any initial obliquity would be washed out already, $\psi_i$ could be neglected in the calculations of tidal heating. Furthermore, the measured eccentricity $e$ could give a constraint to the tidal dissipation function $Q$. Computations based on the theory of `constant-time-lag' yield qualitatively similar results. For the evolution of $e$, we relied on Eq. (\ref{equ:e_FM}). We took the observed eccentricity $e = 0.3216$ as a starting value and evolved it backwards in time. To evolve the system into the past, we changed the sign of the right side of the equation. Furthermore, we assumed that the quality factors $Q_1$ and $Q_2$ of the primary and secondary are equal, leading to $Q_1 = Q_2 \eqqcolon \tilde{Q}$ and $\tilde{\varepsilon}_{k,i} \coloneqq \tilde{\varepsilon}_{k,i}(\Omega_i,n,\tilde{Q})$, because we are merely interested in a tentative estimate so far. This assumption should be a good approximation due to the similarity of the both components in terms of composition, temperature, mass, and radius. The observed eccentricity of the system might give a constraint to the possible values for $\tilde{Q}$ since $\mathrm{d}e/\mathrm{d}t$ depends on $\tilde{Q}$ via $\tilde{\varepsilon}_{k,i}$. Certain $\tilde{Q}$ regimes could be incompatible with the observed eccentricity of the system at a maximum age of 1.5\,Myr, if these $\tilde{Q}$ values would have caused the eccentricity to decay rapidly to 0 within this time. However, our simulations (Fig.~\ref{fig:ecc_I2}) show that the system has not yet evolved very far for the whole range of $\tilde{Q}$ and that the eccentricity of 2M0535$-$05 is in fact increasing nowadays. In this system, circularization does not occur. The observed eccentricity of 0.3216 consequently does not constrain $\tilde{Q}$. In this first estimate, we fixed all other parameters in time, i.e. we neglected an evolution of the semi-major axis $a$, of possible obliquities $\psi_{i}$ and we used constant radii $R_{i}$ and rotational frequencies $\Omega_{i}$. We did this because we cannot yet incorporate the evolutionary behavior of the components' radii $R_{i}$ in the context of tides and furthermore, there is no knowledge about possible misalignments $\psi_{i}$ between the orbital plane and the equatorial planes of the primary and secondary, respectively. A consistent evolution of $R_i$, however, is necessary to evolve $\mathrm{d}a/\mathrm{d}t$ as a function of $\psi_1$ and $\psi_2$, as given by Eq. (\ref{equ:a_FM}). Such a calculation was beyond the scope of this study. The relative spin-geometry of the two BD rotational axes with respect to the orbital plane and with respect to each other is unknown in 2M0535$-$05. Anyhow, we can estimate if a possible obliquity that once existed for one of the BDs would still exist at an age of 1.5\,Myr or if it would have been washed out up to the present. We used a numerical integration of Eq. (\ref{equ:psi_FM}) to evolve $\psi_2$ forward in time (Fig.~\ref{fig:ecc_I2}). For the secondary's initial obliquity $\psi_{\mathrm{ini},2}$, we plot the state of $\psi_2$ as a function of the quality factor $Q_2$ after an evolution of 1.5\,Myr. We see that even for a very small quality factor of $10^3$ and high initial obliquities the secondary is basically in its natal configuration today. Thus, it is reasonable to include a putative misalignment of the secondary with respect to the orbital plane in our considerations. As shown below, this is crucial for the calculations of the tidal heating and the temperature reversal. \subsection{Tidal heating in 2M0535$-$05 with model \#1} \label{sub:tid_mod1_results} In Fig. \ref{fig:E_mod1}, we show the results for the tidal heating rates as computed after tidal model \#1. As given by Eq. (\ref{equ:E_tid_mod1_2M}), the tidal heating of the primary does not depend on a putative obliquity, whereas that of the secondary does. Using this model, we find that the luminosity gain of the secondary is, over the whole $Q$ range, smaller than that of the primary, which mainly results from the relation $\dot{E}_{\mathrm{tid},i}^{\mathrm{\#1}} \propto R_i^5$. Figure \ref{fig:E_mod1} also shows that a growing obliquity shifts the gain in thermal energy towards higher values for a fixed $Q_2$. The observed overshoot of $\approx 10^{24}$\,W in the secondary's luminosity can be reproduced with very small quality factors of $Q_2 \approx 10^3$ and high obliquities up to $\psi_2 \approx 90^{\circ}$. In Fig. \ref{fig:dT_mod1}, we show the results for the temperature increase as per Eq. (\ref{equ:dT2}) with the tidal energy rates coming from model \#1. These rates yield only a slight temperature increase for both constituents. Even for low $Q$ values of order $10^{4}$ and high obliquities of the secondary, the heating only reaches values $\lesssim 10$\,K. We also see that the heating for the primary is computed to be greater than that for the secondary and no temperature reversal would be expected. If both BDs have the same Q values, then model \#1 is unable to explain the temperature reversal. We cannot rule out a system in which, e.g., $Q_1 = 10^5$ and $Q_2 = 10^3$, for which model \#1 could explain the reversal. However, there is no reason to expect that similar bodies have $Q$ values that span orders of magnitude. Hence, we conclude that model \#1 can neither reproduce the luminosity overshoot of the secondary nor the system's temperature reversal. \subsection{Tidal heating in 2M0535$-$05 with model \#2} \label{sub:tid_mod2_results} This model yields the highest heating rates and hence temperature increases. The contrast between the absolute energy rates within the primary $\dot{E}_{\mathrm{tid,1}}^{\mathrm{\#2}}$ and the secondary $\dot{E}_{\mathrm{tid,2}}^{\mathrm{\#2}}$ is very small. In fact, for any given point in $\psi$-$Q$ space, the heating rates differ only by $\log(\dot{E}_{\mathrm{tid,1}}^{\mathrm{\#2}}/\mathrm{W})-\log(\dot{E}_{\mathrm{tid,2}}^{\mathrm{\#2}}/\mathrm{W}) \approx 0.1$ (Fig. \ref{fig:E_mod2}). The tidal energy rates of the secondary become comparable to the observed luminosity overshoot at $\log(Q_2) \approx 3.5$ and $\psi_2 \approx 50^\circ$, where $\dot{E}_{\mathrm{tid,2}}^{\mathrm{\#2}} \approx 10^{24}$\,W. A comparison of the heating rates from model \#2 with those of model \#1 for either of the BDs shows that model \#2 provides higher rates, with growing contrast for increasing obliquities. The temperature increase arising from the comparable heating rates is inverted for a given spot on the $\psi$-$\log(Q)$ plane. If both BDs had the same obliquity and the same dissipation factor, the secondary would experience a higher temperature increase. As presented in Fig. \ref{fig:dT_mod2}, the temperature increase after model \#2 is significant only in the regime of very low $Q$ and high obliquities. Neglecting any orbital or thermal evolution of the system, the observed temperature reversal could be reproduced by assuming an obliquity for the secondary while the primary's rotation axis is nearly aligned with the normal of the orbital plane. We note that the real heating will probably be greater since model \#2 assumes synchronous rotation, which is not the case for both BDs in 2M0535$-$05 (see Table \ref{tab:parameters}). The values of $Q_2$ and $\psi_2$ necessary to account for the observed increase in $L_2$ and $T_{\mathrm{eff},2}$ may thus be further shifted towards more reasonable numbers, i.e. $Q_2$ might also be higher than $10^{3.5}$ and the obliquity might be smaller than $50^\circ$. Thus, for a narrow region in the $\psi$-$\log(Q)$ plane, model \#2 yields tidal energy rates for the secondary comparable to its observed luminosity overshoot and in this region the computed temperature increase can explain the observed temperature reversal. \subsection{Tidal heating in 2M0535$-$05 with model \#3} \label{sub:tid_model_3_results} Since the only free parameter in this model is the putative fixed time lag $\tau$, we show the tidal heating rates for both the primary and the secondary only as a function of $\tau$ in Fig. \ref{fig:E_mod3} with $0\,\mathrm{s} < \tau < 300\,\mathrm{s}$. For this range, model \#3 yields energy rates and temperature rises that are compatible with the observed luminosity and temperature overshoot of the secondary. For $\tau \gtrsim 100$\,s the heating rate for the secondary becomes comparable to the observed one, namely $\dot{E}_{\mathrm{tid},2}^{\mathrm{Hut}} \approx 10^{24}$\,W. However, assuming a similar time lag $\tau_1$ for the primary, the luminosity gain of the primary BD would be significantly higher than that of the secondary, which is not compatible with the observations. The assumption of $\tau_1 \approx \tau_2$ should be valid since both BDs are very similar in their structural properties, such as mass, composition, temperature, and radius. The corresponding temperature increase is plotted in Fig. \ref{fig:dT_mod3}. It shows that the more massive BD would experience a higher temperature increase than its companion, assuming similar time lags. Since tidal heating is underway in 2M0535$-$05 and was probably similar in the past (see Sect. \ref{subsub:tid_model_1}), tidal heating after model \#3 would have been more important on the primary, forcing it to be even hotter than it would be without the perturbations of the secondary. The temperature difference between the primary and the secondary, which is anticipated by BD evolutionary models, would be even larger. Thus, the temperature inversion cannot be explained by tidal model \#3. \subsection{Tidal heating in 2M0535$-$05 with model \#4} \label{sub:tid_model_4_results} The calculations based on model \#4 yield significant heating rates in both BDs. Like in the case of models \#1 and \#2, the luminosity gain of the secondary at a fixed obliquity is, over the whole $Q_{\mathrm{n}}$ range, smaller than that of the primary (Fig. \ref{fig:E_mod4}). As for model \#2, the difference between $\dot{E}_{\mathrm{tid},1}^\mathrm{\#4}(\psi)$ and $\dot{E}_{\mathrm{tid},2}^\mathrm{\#4}(\psi)$ is less pronounced than in model \#1. Assuming spin-orbit alignment for the primary and a pronounced obliquity of the secondary, tidal heating rates of $\dot{E}_{\mathrm{tid},2}^{\mathrm{\#4}} = 10^{24}$\,W can be reached with $\log(Q_{\mathrm{n},2}) \approx 3.5$ and $\psi_2 \approx 50^\circ$. Like model \#2, \#4 produces a reversal in temperature increase by means of the modified Stefan-Boltzmann relation in Eq. (\ref{equ:dT2}), due to the comparable heating rates of both BDs and the significantly smaller radius of the secondary (Fig. \ref{fig:dT_mod4}). We find a reversal in tidal heating, i.e. $\mathrm{d}T_2 > \mathrm{d}T_1$ for any given point in $\psi$-$Q_{\mathrm{n}}$ space. A temperature increase of $\gtrsim 40$\,K can be reached with $\log(Q_{\mathrm{n},2}) \approx 3.5$ and $\psi_2 \approx 50^\circ$. Since the equations of model \#4 provide merely a lower limit due to the assumption of asymptotic non-synchronous rotation, $Q_{\mathrm{n},2}$ might also be higher than $10^{3.5}$ and the obliquity might be smaller than $50^\circ$. Similar to model \#2, tidal model \#4 can reproduce the observed temperature reversal in a narrow region of the $\psi$-$\log(Q)$ parameter space. \section{Discussion} \label{sec:discussion} We employed several tidal models to explore the tidal heating in 2M0535$-$05. We found that, assuming similar tidal quality factors $Q$ and obliquities $\psi$ for both BDs, the constant-phase-lag model \#2 and the constant-time-lag model \#4 yield a stronger increase in effective temperature on the secondary mass BD than on the primary. For certain regimes of $Q_2$ and $\psi$, the tidal energy rates in the secondary are of the correct amount to explain the larger temperature in the smaller BD. A comparison between our computations based on the models \#1 and \#2 on the one hand and \#3 and \#4 on the other hand is difficult. The reference to a fixed tidal time lag might only be reconciled with the assumption of $Q_{\mathrm{n}}^{-1} = n \ \tau$ as done by Lev07, which is at least questionable since the assumption of a fixed time lag is not compatible with a fixed phase lag. Furthermore, model \#3 does not invoke obliquities, which also complicates direct comparisons of the model output. \subsection{Constraints on the tidal dissipation function for BDs, $Q_{\mathrm{BD}}$} \label{sec:constraints_Q} \subsubsection{The Rossiter-McLaughlin effect in 2M0535$-$05} \label{subsub:rme} The geometric implication of the most promising tidal models \#2 and \#4 is that the obliquity of the 2M0535$-$05 primary is negligible and that of the secondary is $\psi_2 \approx 50^\circ$ -- provided tidal heating accounts for the $T_{\mathrm{eff}}$ reversal and the luminosity excess of the secondary. There does exist an observational method to measure the geometric configuration of eclipsing systems, called the Rossiter-McLaughlin effect (RME) \citep{1924ApJ....60...15R, 1924ApJ....60...22M}. The RME appears during transits in front of rotating stars. Hiding a fraction of the star's surface results in the absence of some corresponding rotational velocity contribution to the broadening of the stellar lines. Thus, the changes in the line profiles become asymmetric (except for the midpoint of the transit) and the center of a certain stellar line is shifted during a transit, which induces a change of the star's radial velocity. The shape of the resulting radial velocity curve depends on the effective area covered by the transiting object and its projected path over the stellar surface with respect to the spin axis of the covered object \citep[for a detailed analysis of the RME see][]{2005ApJ...622.1118O}. Using a code originally presented in \citet{2009A&A...499..615D}, we have undertaken simulations of the RME for various geometric configurations of 2M0535$-$05 during the primary eclipse\footnote{The `primary eclipse' refers to the major flux decrease in the system's light curve. Due to the significantly higher effective temperature of the secondary mass BD the primary eclipse occurs when the primary mass component transits in front of the secondary companion, as seen from Earth.} as it would be seen with the Ultraviolet and Visual Echelle Spectrograph (UVES) at the Very Large Telescope (VLT) (see Fig.~\ref{fig:RM}). For the data quality we assumed the constraints given by the UVES at the VLT exposure time calculator\footnote{http://www.eso.org/observing/etc} in version 3.2.2. The computations show that, using Th-Ar reference spectra and also the telluric A and B bands as benchmarks, a time sampling with one spectrum every 1\,245\,s and a S/N of $\gtrsim$ 7 around 8\,600\,$\AA$ are necessary to get 21 measurements during the primary eclipse and an accuracy of $\lesssim$ 100\,m/s. In principle, there are four parameters for the background object of the transit to be fitted in our simulations of the RME: the rotational velocity $v_{\mathrm{rot}}$, the inclination of the spin axes with respect to the line of sight $I_{\star}$, the angle between the projection of the spin and the projection of the orbital plane normal onto the celestial plane $\lambda$, and the orbital inclination with respect to the line of sight $i$. From light curve analyses, both rotational velocities in 2M0535$-$05 and the orbital inclination $i$ are known. Thus, for the simulation of the primary eclipse $I_{\star,2}$ and $\lambda_2$ are the remaining free parameters. The obliquities $\psi_{i \ | \ i = 1, 2}$, i.e. the real 3-dimensional angle between the orbital normal and the spin axis of the occulted object, is related to the other angles as \begin{equation}\label{equ:psi_definition} \cos(\psi_i) = \cos(I_{\star,i}) \cos(i) + \sin(I_{\star,i}) \sin(i) \cos(\lambda_i) . \end{equation} \noindent While the two obliquities $\psi_i$ are intrinsic angles of the system, they cannot be measured directly. They can only be inferred from $i$, $I_{\star,i}$ and $\lambda_i$, which depend on the position of the observer with respect to the system. Since we are only interested in the possible options for the measurement of the obliquities in 2M0535$-$05, we refer the reader to the paper by \citet{2005ApJ...631.1215W} for a discussion of Eq. (\ref{equ:psi_definition}) and the geometrical aspects of the RME. With $i = 88.49^{\circ}$ the first term in Eq. (\ref{equ:psi_definition}) degrades to insignificance, which yields $\cos(\psi_i) \approx \sin(I_{\star,i}) \cos(\lambda_i)$. At low values for $I_{\star,i}$ and $\lambda_i$ the fitted solutions to the RME are degenerate and there are multiple solutions within a certain confidence interval. But our simulations for the transit show that the error due to the observational noise is on the same order as the error due to degeneracy and thus we find standard deviations in $I_{\star,2}$ and $\lambda_2$ of $\sigma_{I_{\star,2}} \approx 20^{\circ}$ and $\sigma_{\lambda_{2}} \approx 20^{\circ}$, respectively. The uncertainty in $\psi_2$ depends not only on the uncertainties in $I_{\star,2}$ and $\lambda_{2}$ but also on the actual values of $I_{\star,2}$ and $\lambda_{2}$. But in all cases, the standard deviation in the secondary's obliquity $\sigma_{\psi_2}~<~20^\circ$. If present in 2M0535$-$05, a considerable misalignment of the secondary BD of $50^\circ$ could be detected with a 1-$\sigma$ accuracy of $20^\circ$ or less. Thus, an observed $\psi_2$ value of $50^\circ$ would be a 2.5-$\sigma$ detection of spin-orbit misalignment. Unless RME measurements suggest $\psi \approx 90^\circ$, RME observations alone are unlikely to provide definitive evidence that any of the tidal models we consider is responsible for the temperature reversal. \subsubsection{Further observations of BD binaries} \label{subsub:further} Besides the option of RME measurements for testing the geometric implications, there does exist a possibility to verify our estimate of $\log(Q) \approx 3.5$ for BDs in general. Comparison of observed orbital properties with values constrained by the equations that govern the orbital evolution might constrain the free parameters, here $Q$. Using Eq. (\ref{equ:e_FM}), we find that, assuming only a slight initial eccentricity of 0.05, the eccentricity of a BD binary system similar to 2M0535$-$05, in terms of masses, radii, rotational frequencies, and semi-major axis would increase to 1 after $\approx 500$\,Myr if the quality factors of the two BDs are $\lesssim 10^{3.5}$ (see left panel in Fig.~\ref{fig:ecc_500and100Myr}). A measurement of $e$ in such an evolved state could not constrain $Q$ in a 2M0535$-$05 analog since either the initial eccentricity could have been relatively large while the orbit evolved rather slowly due to high $Q$ values or a small initial value of $e$ could have developed to a large eccentricity due to small values of $Q$. We also simulate the evolution of a 2M0535$-$05 analog but with a different rotational frequency of the primary constituent in order to let the eccentricity decrease with time. We neglected the evolution of all the other physical and orbital parameters since we are merely interested in a tentative estimate. For a given candidate system the analysis would require a self-consistent coupled evolution of all the differential equations. For the arbitrary case of $P_1 = P_2 = 14.05$\,d we find that, even for the most extreme but unrealistic case of an initial eccentricity equal to 1, this fictitious binary would be circularized on a timescale of 100\,Myr for $\log(\tilde{Q}) < 5$ (see right panel in Fig.~\ref{fig:ecc_500and100Myr}). Findings of old, eccentric BD binaries with rotational and orbital frequencies that yield circularization in the respective system would set lower limits to $Q$. \subsubsection{Rotational periods in 2M0535$-$05} \label{subsub:rotation} Another, and in fact a crucial, constraint on $Q$ for BDs comes from the synchronization time scale $t_\mathrm{synch}$ of the two BDs in 2M0535$-$05. Following the equation given in Lev07 and taking the initial orbital mean motion and semi-major axis of the system as calculated with an uncoupled system of differential equations from model \#1, we derive $t_{\mathrm{synch},1} = 0.07$\,Myr for the primary and $t_{\mathrm{synch},2} = 0.04$\,Myr for the secondary with $\log(Q) = 3.5$. Since the rotation in both BDs is not yet synchronized with the orbit and the age of the system is about 1\,Myr, $\log(Q) = 3.5$ is not consistent with the age of 2M0535$-$05. Both components should have synchronous rotation rates already. We find the consistent value for $Q$ to be $\gtrsim 10^{4.5}$, yielding synchronization time scales $t_{\mathrm{synch},1} \gtrsim 0.69$\,Myr and $t_{\mathrm{synch},2} \gtrsim 0.37$\,Myr. To make this estimate for $Q$ more robust, we present the evolution of the BDs' rotational periods in Fig.~\ref{fig:rot_evol} and compare it to the critical period for a structural breakup $P_\mathrm{crit}$. The evolutionary tracks are calculated with model \#1 and Eq. (30) in FM08. As a rough approach we do not couple this equation with those for the other orbital parameters. The left panel of Fig.~\ref{fig:rot_evol} shows that for $\log(Q_1) = 3.5$ and $\psi_1 = 0^\circ$ the primary's initial rotation period 1\,Myr ago is $\approx 0.3$\,d. The initial rotation period for the secondary, for $\log(Q_2) = 3.5$ and $\psi_2 = 0^\circ$, is about -0.2\,d, where the algebraic sign contributes for a retrograde revolution (right panel in Fig.~\ref{fig:rot_evol}). For most of its lifetime, the secondary would have had a retrograde rotation and just switched the rotation direction within the last few 10,000 yr, which is very unlikely in statistical terms. Since the orbital momentum is on the order of $10^{43}$\,kg\,m$^2$\,/\,s and the individual angular momenta are about $10^{41}$\,kg\,m$^2$\,/\,s, the shrinking process might not have had a serious impact on the rotational evolution. Tides have dominated the spin evolutions. Following \citet{2005A&A...429.1007S}, the critical breakup period $P_\mathrm{crit}$ depends only on the body's radius and its mass. The radius evolution for BDs is very uncertain for the first Myr after formation but we estimate their initial radii to be as large as the solar radius. This yields $P_{\mathrm{crit},1} \approx 0.5$\,d for both the primary and the secondary BD. As stated above, the moduli of the initial rotation periods of both BDs would have been smaller than 0.5\,d for $Q$ values of $\lesssim 10^{3.5}$. This inconsistency gives a lower limit to $Q_1$ and $Q_2$ since values of $\lesssim 10^{3.5}$ would need an initial rotation periods of both BDs which are smaller than their critical breakup periods. Obliquities larger than $0^\circ$ would accelerate the (backwards) evolution and yield even larger lower limits for $Q_1$ and $Q_2$. Thus, our simulations of the rotational period evolution of both BDs require $\log(Q_{\mathrm{BD}}) \gtrsim 3.5$, whereas the tidal synchronization timescale even claims $\log(Q_{\mathrm{BD}}) \gtrsim 4.5$. \subsection{Evolutionary embedment of tidal heating} \label{sub:embedment} Tidal heating must be seen in the evolutionary context of the system. On the one hand, the tidal energy rates generate a temperature increase on the Kelvin-Helmholtz time-scale, which is $\approx 2$\,Myr for the BDs in 2M0535$-$05 -- and thus on the order of the system's age, as per Eq. (\ref{equ:dT2}). On the other hand, tidal heating will affect the shrinking and cooling process of young BDs in terms of an evolutionary retardation. As models show \citep{1997MmSAI..68..807D, 1998A&A...337..403B, 2000ApJ...542..464C, 2000ARA&A..38..337C}, single BDs cool and shrink significantly during their first Myrs after formation. Adding an energy source comparable to the luminosity of the object will slow down the aging processes such that the observed temperature and luminosity overshoot at some later point is not only due to the immediate tidal heating but also due to its past evolution. Consequently, the luminosity and temperature overshoot in the secondary might not (only) be due to present-day tidal heating, but it could be a result of an evolutionary retardation process triggered by the presence of the primary as a perturber. Coupled radius-orbit evolutionary models have already given plausible explanations for the inflated radii of some extrasolar planets \citep{2003ApJ...588..509G, 2009ApJ...702.1413M, 2009ApJ...700.1921I, 2009arXiv0910.4394I, 2009arXiv0910.5928I}. For a consistent description of the orbital and physical history of 2M0535$-$05, one would have to include the evolution of obliquities $\psi_i$, BD radii $R_i$, eccentricity $e$, semi-major axis $a$, and rotational frequencies $\Omega_i$. Note that there is a positive feedback between radial inflation and tidal heating: as tidal heating inflates the radius, the tidal heating rate can increase and -- in turn -- may cause the radius to inflate even more. In a self-consistent orbital and structural simulation of 2M0535$-$05, tidal inflation, neglected in our computations of the $T_{\mathrm{eff}}$ increase in Eq. (\ref{equ:dT2}), will result naturally from the additional heating term introduced by tides. In conjunction with 2M0535$-$05 that means the actual heating rates necessary to explain the $T_{\mathrm{eff}}$ and luminosity excess in the secondary are lower than they would have to be if there would be no historical context. Relating to Figs. \ref{fig:dT_mod1}, \ref{fig:dT_mod2}, \ref{fig:dT_mod3}, and \ref{fig:dT_mod4}, the implied obliquity and $Q$ factor for the secondary are -- again -- shifted towards lower and higher values, respectively. Embedded in the historical context of tidal interaction in 2M0535$-$05, $\psi_2 < 50^{\circ}$ and $\log(Q_{2}) >3.5$ may also explain the temperature reversal and the luminosity excess of the secondary. These trends, however, are contrary to that induced by tidal inflation. If tidal heating is responsible for a radial expansion of 10 and 20\% in the primary and secondary, the values of the dissipation factor necessary to explain the $T_\mathrm{eff}$ reversal would be $\approx 0.8$ smaller in $\log(Q_2)$ (see Sect. \ref{sub:converting}). \section{Conclusions} \label{sec:conclusions} We surveyed four different published tidal models, but neglect any evolutionary background of the system's orbits and the components' radii to calculate the tidal heating in 2M0535$-$05. Our calculations based on models \#2 and \#4, which are most compatible with the observed properties of the system, require obliquities $\psi_1 \approx 0$, $\psi_2 \approx 50^{\circ}$ and a quality factor $\log(Q) \approx 3.5$ in order to explain the luminosity excess of the secondary. Additionally, the observed temperature reversal follows naturally since we may reproduce a reversal in temperature increase due to tides: $\mathrm{d}T_2 > \mathrm{d}T_1$. In model \#2, synchronous rotation of the perturbed body is assumed. Since this is not given in 2M0535$-$05, the actual heating rates will be even higher than those computed here. Our results for the heating rates as per model \#2 are thus lower limits, which shifts the implied obliquity of the secondary and its $Q$ factor to lower and higher values, respectively. Considerations of the synchronization time scale for the BD duet and the individual rotational breakup periods yield constraints on $Q_{\mathrm{BD}}$ for BDs. We derive a lower limit of $\log(Q_{\mathrm{BD}}) > 4.5$. This is consistent with estimates of $Q$-values for M dwarfs, $\log(Q_{\mathrm{dM}}) \approx 5$, and the quality factors of Jupiter, $2 \cdot 10^{5} < Q_{\jupiter} < 2 \cdot 10^{6}$, and Neptune, $10^{4} \lesssim \log(Q_{\neptune}) \lesssim 10^{4.5}$ (see Sect. \ref{subsub:tid_model_1}). With $\log(Q_{\mathrm{BD}}) > 4.5$ tidal heating alone can neither explain the temperature reversal in the system nor the luminosity excess of the secondary. An obliquity of $50^\circ$, however, would be reasonable in view of recent results from measurements of the RME in several transiting exoplanet systems\footnote{See http://www.hs.uni-hamburg.de/EN/Ins/Per/Heller for an overview.}. Currently, out of 18 planets there are 7 with significant spin-orbit misalignments $\gtrsim 30^\circ$ and some of them are even in retrograde orbits around their host stars. A substantial obliquity $\psi_2$ might cause an enhanced heating in the 2M0535$-$05 secondary, while the primary's spin could be aligned with the orbital spin, leading to negligible heating in the primary. Despite the advantages of distance-independent radius and luminosity measurements of close, low-mass binaries, the comparison of fundamental properties of the constituents with theoretical models of isolated BDs must be taken with care. This applies also to the direct translation from the discrepancies between observed and modeled radii for a fixed metallicity into an apparent age difference as a calibration of LMS models \citep{2009IAUS..258..161S}. Tidal heating might be a crucial contribution to discrepancies between predicted and observed radii in other eclipsing low-mass binary systems \citep{2008MmSAI..79..562R}. As recently shown by \citet{2009ApJ...700.1921I}, tidal heating in extra-solar giant planets in close orbits at $a \lesssim 0.2$\,AU with modest to high eccentricities of $e \gtrsim 0.2$ can explain the increased radii of some planets, when embedded in the orbital history with its host star. Improvement of tidal theories is necessary to estimate the relation between tides and the observed radii of LMS being usually too large as compared to models. A tidal model is needed for higher orders of arbitrary obliquities and eccentricities that also accounts for arbitrary rotation rates. As stated by \citet{2009ApJ...698L..42G}, a formal extension of the simple `lag-and-add' procedure of tidal frequencies the theory of constant phase lag is questionable. Besides the extension, conciliation among the various models is needed. The results from the models applied here should be considered preliminary but are suggestive and indicate the possible importance of tides in binary BD systems. Several issues remain to be addressed for a more detailed assessment of tidal heating in 2M0535$-$05: \textit{i.} reconciliation and improvement of tidal theories; \textit{ii.} self-consistent simulations of the orbital and physical evolution of the system and the BDs; \textit{iii.} measurements of the system's geometric configuration; \textit{iv.} constraints on the tidal quality factors of BDs. \begin{acknowledgements} Our sincere thanks go to J. L. Bean for initiating this collaboration. The advice of S. Dreizler on the computations of the RM effect and inspirations from A. Reiners on 2M0535$-$05 and BDs in general have been a valuable stimulation to this study. We acknowledge the help of Y. G. M. Chew and K. G. Stassun on the parametrization of the 2M0535$-$05 BD binary and we appreciate the contribution from I. Baraffe to the modeling of the BDs' structures. The referee Jean-Paul Zahn deserves our honest gratitude for his crucial remark on the tidal synchronization time scale. R. Heller is supported by a PhD scholarship of the DFG Graduiertenkolleg 1351 ``Extrasolar Planets and their Host Stars''. R. Barnes acknowledges funding from NASA Astrobiology Institute's Virtual Planetary Laboratory lead team, supported by NASA under Cooperative Agreement No. NNH05ZDA001C. R. Greenberg, B. Jackson, and R. Barnes were also supported by a grant from NASA's Planetary Geology and Geophysics program. This research has made use of NASA's Astrophysics Data System Bibliographic Services. \end{acknowledgements} \bibliographystyle{aa}
\section{Introduction} The starting point of this work is the following ``Contraction Theorem" of L. Caffarelli \cite{CaffarelliContraction}: \begin{thm*}[Caffarelli] Let $\mu = \exp(-Q(x)) dx$ and $\nu = \exp(-(Q(x) + V(x))) dx$ denote two Borel probability measures on Euclidean space $(\Real^n,\abs{\cdot})$, where $Q$ denotes a quadratic function, i.e. \begin{equation} \label{eq:quad} Q(x) = \scalar{Ax,x} + \scalar{b,x} + c ~, \end{equation} with $A$ positive-definite, and $V$ is a convex function. Then the Brenier optimal-transport map $T = T_{opt}$ pushing forward $\mu$ onto $\nu$ is a contraction: \[ \forall x,y \in \Real^n \;\;\; |T(x) - T(y)| \leq |x - y| ~. \] \end{thm*} Let us recall some of the notions used above. A Borel map $T$ is said to push-forward $\mu$ onto $\nu$, denoted $T_*(\mu) = \nu$, if $\nu(A) = \mu(T^{-1}(A))$ for any Borel set $A$. Among all such maps $T$, it is natural to minimize the squared-distance transport cost: $W_2^2(\mu,\nu):= \inf_{T_*(\mu) = \nu} \int |T(x) - x|^2 d\mu(x)$ - this is precisely the Monge (or Monge-Kantorovich) problem for a quadratic cost. The Brenier map $T_{opt} : \Real^n \rightarrow \Real^n$ pushing forward $\mu$ onto $\nu$ is the $\mu$-a.e. unique map for which the latter infimum is attained; it is precisely characterized by the property of being the gradient of a convex function $\varphi: \Real^n \rightarrow \Real$, as first proved by Y. Brenier \cite{BrenierMap}. It is known that the optimal-transport distance $W_2$ metrizes the Wasserstein space $W_2(\Real^n)$ of square integrable Borel probability measures on $\Real^n$ equipped with a suitable weak topology. We refer to \cite{VillaniTopicsInOptimalTransport,VillaniOldAndNew} for a comprehensive account on this and related topics. \subsection{Main Result} Fix an orthogonal decomposition of $(\Real^n,|\cdot|)$ into subspaces $\set{E_i}_{i=0}^k$. \begin{dfn*} We will say that a function $F: \Real^n \rightarrow \Real$ \emph{satisfies our symmetry assumptions} if it is invariant under the action of the subgroup $O(E_1,\ldots,E_k) := 1 \times O(E_1) \times \ldots O(E_k)$ of the orthogonal group $O(n)$, or equivalently, if: \begin{equation} \label{eq:symmetry} \text{$\exists \, \Phi : \Real^{dim E_0 + k} \rightarrow \Real \;\;\text{so that}\;\; F(x) = \Phi(Proj_{E_0} x,|Proj_{E_1} x|,\ldots,|Proj_{E_k} x|)$} ~. \end{equation} We will similarly say that a map $T : \Real^n \rightarrow \Real^n$ \emph{satisfies our symmetry assumptions} if it commutes with the action of the latter subgroup. \end{dfn*} Our main result generalizes Caffarelli's Theorem as follows: \begin{thm} \label{thm:main1} Let $\mu = \exp(-U(x)) dx$ and $\nu = \exp(-(U(x) + V(x))) dx$, denote two Borel probability measures on Euclidean space $(\Real^n,\abs{\cdot})$. Assume that $U \in C^{3,\alpha}_{loc}(\Real^n)$ ($\alpha > 0$) is a convex function of the form: \begin{equation} \label{eq:U} U(x) = Q(Proj_{E_0} x) + \sum_{i=1}^k \rho_i(|Proj_{E_i} x|) ~,~ \forall i = 1 \ldots k ~ ~ \rho'''_i \leq 0 \text{ on $\Real_+$ }~, \end{equation} where $Q : E_0 \rightarrow \Real$ is a quadratic function as in (\ref{eq:quad}), and that $V : \Real^n \rightarrow \Real$ is convex and satisfies our symmetry assumptions (\ref{eq:symmetry}). Then there exists a map $T: \Real^n \rightarrow \Real^n$ pushing forward $\mu$ onto $\nu$ and satisfying our symmetry assumptions which is a contraction. \end{thm} \begin{rmk*} The smoothness assumption on $U$ above is immaterial, and may be dispensed if $\mu$ is approximated (say, in total-variation distance) by measures $\set{\mu_l}$ which satisfy the conditions of Theorem \ref{thm:main1} (see Lemma \ref{lem:approx}). A prototypical example where this applies is for the functions $\rho_i(x) = |x|^{p_i}$, $p_i \in [1,2]$. The same comment applies for $V$, so it is enough to prove the theorem for smooth $U,V$, and conclude by a compactness argument detailed in Section \ref{sec3}. \end{rmk*} \medskip The general formulation of Theorem \ref{thm:main1} interpolates between the following extremal cases: \begin{itemize} \item $\mu$ is a product measure and $V$ is ``unconditional": \[ \text{$U(x) = \sum_{i=1}^n \rho_i(|x_i|)$ with $\rho_i''' \leq 0$ and $V(x_1,\ldots,x_n) = V(\pm x_1,\ldots,\pm x_n)$ are convex} ~. \] \item $U$ and $V$ are both radial: \[ \text{$U(x) = \rho(|x|)$ with $\rho''' \leq 0$ and $V(x) = \Phi(|x|)$ are convex} ~. \] \item $U$ is quadratic and $V$ is an arbitrary convex function. \end{itemize} We shall be mainly interested in the first case, since the third one follows immediately from Caffarelli's result, and the second one may be easily obtained using a one dimensional argument reproducing Caffarelli's original proof, as described in Section \ref{sec:revisit}. However, for some of the applications presented in this work, the case when $0 < dim E_0 < n$ is the most interesting. We also remark that Caffarelli's theorem has recently been generalized in other directions by Valdimarsson \cite{ValdimarssonGeneralizedCaffarreli} and Kolesnikov \cite{KolesnikovGeneralizedCaffarelli}. \subsection{The Construction}\label{SS:construction of T} As opposed to the non-constructive optimal-transport map $T_{opt}$, our map $T$ is obtained as a limit of diffeomorphisms $\set{T_t}_{t \geq 0}$, constructed as a (reverse) flow along an advection field generated by an appropriate heat diffusion process. Let $L$ denote the following second-order differential operator: \begin{equation} \label{eq:L-def} L = \exp(U) \; \nabla \cdot (\exp(-U) \nabla) = \Delta - \scalar{\nabla,\nabla U} ~, \end{equation} and let $P^U_t := \exp(t L) : L_\infty(\Real^n) \rightarrow L_\infty(\Real^n)$ denote the associated diffusion semi-group, characterized as solving the parabolic equation: \begin{equation} \label{eq:semi-group} \frac{d}{dt} P^U_t(f) = L(P^U_t(f)) ~,~ P^U_0(f) = f ~ ~ \text{(for smooth bounded functions $f$)} ~. \end{equation} The latter is simply the usual heat-equation with an additional first-order drift term, also known as the (linear) Fokker-Planck equation. Its invariant measure is easily checked to be $\mu = \exp(-U(x)) dx$: \begin{equation} \label{eq:L-prop} \int L(f) g d\mu = - \int \scalar{\nabla f, \nabla g} d\mu = \int f L(g) d\mu ~,~ \int P^U_t(f) g d\mu = \int f P^U_t(g) d\mu ~. \end{equation} In particular, $-L$ becomes a self-adjoint positive semi-definite operator on an appropriate dense subspace of $L_2(\mu)$. Since $\mu$ is a log-concave probability measure, it is known that $-L$ has a non-trivial spectral-gap, from which it follows by the Spectral Theorem that $P_t^U(f) \rightarrow_{t \rightarrow \infty} \int f d\mu$ in a rather strong sense (see Section \ref{sec3}). Defining: \begin{equation} \label{eq:nu-t-def} \nu_t := P^U_t(\exp(-V)) \mu ~, \end{equation} it follows in particular that $\nu_0 = \nu$ and $\nu_t \rightarrow_{t \rightarrow \infty} \mu$, so $\set{\nu_t}$ naturally interpolate between $\nu$ and $\mu$. We will show how to construct diffeomorphisms $\set{T_t}_{t \geq 0}$, so that each $T_t$ is a contraction satisfying our symmetry assumptions which pushes forward $\nu_t$ onto $\nu$. Theorem \ref{thm:main1} then follows by a compactness argument, ensuring that $\set{T_t}$ converge appropriately to our desired map $T$. Our construction is in fact for the inverse-maps $S_t := T_t^{-1}$, pushing forward $\nu$ onto $\nu_t$. These diffeomorphisms are constructed as a flow along a (time-dependent) advection field $W_t$ induced by our diffusion: \begin{equation} \label{eq:vector-flow} \frac{d}{dt} S_t(x) = W_t(S_t(x)) ~,~ S_0 = Id ~. \end{equation} To choose a consistent $W_t$, we use the well-known Continuity Equation (see e.g. \cite{VillaniTopicsInOptimalTransport}): \[ \frac{d}{dt} \nu_t + \nabla \cdot (\nu_t W_t) = 0 ~, \] which allows us to pass from the Lagrangian view point (\ref{eq:vector-flow}) to an Eulerian one. We conclude using (\ref{eq:nu-t-def}) that: \[ \frac{d}{dt} P^U_t(\exp(-V)) = - \exp(U) \; \nabla \cdot (\exp(-U) P^U_t(\exp(-V)) W_t ) ~, \] and to make this consistent with (\ref{eq:L-def}) and (\ref{eq:semi-group}), we choose: \begin{equation} \label{eq:W-t-def} W_t := - \nabla \log P^U_t(\exp(-V)) ~. \end{equation} It remains to show that the maps $S_t$ are expansions, i.e. $|S_t(x) - S_t(y)| \geq |x-y|$. Being diffeomorphisms, this is equivalent to requiring that the maps are expansions \emph{locally}: \[ (DS_t)^* DS_t \geq Id ~. \] Differentiating this inequality in $t$ and using (\ref{eq:vector-flow}), we see that it suffices to show that $D W_t + (D W_t)^*\geq 0$ for all $t \geq 0$. By (\ref{eq:W-t-def}), this is equivalent to showing that: \[ - D^2 \log P^U_t(\exp(-V)) \geq 0 \;\;\; \forall t \geq 0 ~. \] \subsection{The Reduction} This is formulated in the following result, which we believe is of independent interest: \begin{thm} \label{thm:main2} Under the assumptions of Theorem \ref{thm:main1}, $P^U_t$ preserves the log-concavity of $\exp(-V)$. In other words, $-\log P^U_t(\exp(-V))$ is a convex function for all $t \geq 0$. \end{thm} It should be noted that by a result of A. Kolesnikov \cite{KolesnikovPreservingLogConcavity} (see also \cite{LionsMusielaPreservingConvexityCRAS}, and compare with \cite{IshigeSalaniQuasiConcavityIsNotPreserved}), the only smooth linear diffusion processes (\ref{eq:semi-group}) with generator $L = A(x) \nabla^2 + b(x) \nabla$ which preserve the log-concavity of $\exp(-V)$ for \emph{arbitrary} convex functions $V$, are precisely the Ornstein-Uhlenbeck processes, given by a constant valued matrix $A$ and an affine map $b$ (for our generator (\ref{eq:L-def}), this corresponds to quadratic potentials $U=Q$). That the Ornstein-Uhlenbeck processes preserve log-concavity is well known, and may be easily seen using the Mehler formula and the Prekop\'a-Leindler Theorem (e.g. \cite{HargeGCCForEllipsoid}); together with our construction above, this already provides an alternative proof of Caffarelli's Contraction Theorem (with some other map $T$). By \emph{restricting} to convex functions $V$ having certain symmetries, as in Theorem \ref{thm:main2}, we are able to show that log-concavity is preserved for generators with \emph{more general} potentials $U$. The proof of Theorem \ref{thm:main2} is based on parabolic PDE methods and in particular the maximum principle (see \cite{KawohlBook,LeeVazquezParabolicNonlinearPDE,GrecoKawohl} and the references therein). Let us give a very heuristic outline of the proof. After assuming that $V$ is smooth enough and strictly convex, and restricting the problem onto a smooth, bounded and strictly convex domain by imposing zero Dirichlet boundary conditions, we proceed in the contrapositive. Assume that $V = V(x,t)$ does not remain strictly convex, and argue that there will be a first time $t_0 > 0$ when this fails; this step is the most delicate in all of the proof and requires very careful justification, a point that has been omitted in many previous works on concavity properties of solutions to parabolic PDE. The strict convexity of the boundary guarantees that the minimum of $D^2_{e,e}V(x,t_0)$ will be attained in an interior point $x_0$ and some direction $e$. Since this will be a local minimum, this implies on one hand that $(d/dt - \Delta)(D^2_{e,e}V)(x_0,t_0) \leq 0$. On the other hand, using that $D D_e V = 0$ and $D D^2_{e,e} V = 0$ at $(x_0,t_0)$, a calculation shows that: \[ \left . \brac{(d/dt - \Delta)(D^2_{e,e}V)} \right |_{(x_0,t_0)} = - \left (D^3 U) \right|_{x_0} (e,e,\grad V(x_0,t_0)) ~. \] At time $t=t_0$, $V(\cdot,t)$ is still assumed to be convex, and our geometric structural and symmetry assumptions on $U$ and $V$ were precisely designed to guarantee that the latter expression be non-negative. Massaging this argument a little more, we obtain a contradiction, thereby concluding the proof. We emphasize again that key to our approach is an analysis at the very first time $t_0>0$ when things may go wrong - a triviality for the usual application of the maximum principle for a (uniformly continuous) function on a bounded parabolic domain, but a genuine issue when applied to its second derivatives, which may not be uniformly continuous up to the boundary. \subsection{Applications} Besides the applications provided in his original paper \cite{CaffarelliContraction}, Caffarelli's Contraction Theorem has found numerous applications in various fields, serving as a tool to transfer isoperimetric inequalities, obtaining correlation inequalities, and more (see e.g. \cite{CorderoMassTransportAndGaussianInqs,CFM-BConjecture,HargeCorrelationInqs,KlartagMarginalsOfInequalities}). Most of these applications only use the fact that there exists \emph{some} contracting map pushing forward one measure onto another, without employing the additional information that this map is the \emph{Brenier} map, i.e. the gradient of a convex function. Consequently, it is a mere exercise to repeat the corresponding proofs in our more general setting, replacing Caffarelli's Theorem with Theorem \ref{thm:main1}, and thereby extending these applications. We will not go through all of these in this work, but rather indicate several selected applications pertaining to correlation inequalities, extending in particular some known results regarding the Gaussian Correlation Conjecture (described in Section \ref{sec:apps}) to our setup, following an argument of Dario Cordero-Erausquin \cite{CorderoMassTransportAndGaussianInqs}. We will also briefly indicate how to obtain new isoperimetric inequalities. \subsection{Afterthoughts} After understanding how to extend Caffarelli's Contraction Theorem using our heat-induced flow and proving Theorem \ref{thm:main1}, we revisited Caffarelli's original argument from \cite{CaffarelliContraction}, and observed: \begin{thm} \label{thm:main3} Theorem \ref{thm:main1} is also valid when replacing $T$ with the Brenier optimal-transport map $T_{opt}$ pushing forward $\mu$ onto $\nu$. \end{thm} For the proof of Theorem \ref{thm:main3}, which is based on Caffarelli's own proof, we require an additional ingredient from \cite{CaffarelliContraction} in the form of Theorem \ref{thm:ingred}, described in Section \ref{sec:revisit}. Roughly speaking, Caffarelli's argument is oblivious to the quadratic part of $U$, and for the non-quadratic part on $E_0^\perp$, reduces under our assumptions the task of showing that $T_{opt}$ is a contraction, to showing that it is a contraction \emph{with respect to the origin}. It is this latter property which is verified using Theorem \ref{thm:ingred}. \medskip In Section \ref{sec:comparing}, we compare between the two maps $T$ (as constructed in Subsection~\ref{SS:construction of T}) and $T_{opt}$. It is not hard to verify that the path $[0,\infty) \ni t \mapsto S_t$ of our interpolating diffeomorphisms does not coincide in general with the path $[0,1) \ni s \mapsto (1-s) Id + s S_{opt}$ of optimal interpolating maps, where $S_{opt} = T_{opt}^{-1}$ denotes the Brenier map pushing forward $\nu$ onto $\mu$. Indeed, our diffusion process may be seen as the gradient flow for the entropy functional $H(\nu_t | \mu)$ on the Wasserstein space $W_2(\Real^n)$ equipped with an appropriate Riemannian structure (Otto and Villani \cite{OttoVillaniHWI}, see also Jordan--Kinderlehrer--Otto \cite{JKO-FokkerPlanckAsGradientDescent}); optimal-transport, on the other hand, corresponds to moving along the geodesic between $\nu$ and $\mu$ in $W_2(\Real^n)$, i.e. gradient flow for the distance squared functional $W_2^2(\nu_t ,\mu)$. Consequently, we believe that the limiting maps $T$ and $T_{opt}$ are in general different, although we have not been able to exclude the possibility that they coincide. The assumptions of Theorem \ref{thm:main1} were precisely designed to ensure that $T$ contracts distances, but it is quite surprising that exactly the same assumptions imply (for seemingly different reasons!) the same for $T_{opt}$. When comparing these two approaches, it is worth pointing out that our diffusion approach only relies on classical regularity results for linear parabolic PDEs, whereas analyzing the optimal-transport map requires Caffarelli's deeper regularity results for the fully-nonlinear Monge-Amp\`ere equation (see \cite{CaffarelliStrictlyConvexIsHolder,CaffarelliRegularity} and the references therein); consequently, the former approach may lend itself to further generalization, in particular to setups where the latter regularity results for the Brenier--McCann optimal-transport map are unavailable, or alternatively, known to be false, as in the Riemannian-manifold setting (see \cite{VillaniOldAndNew}). \subsection{Organization} The rest of this work is organized as follows. In Section \ref{sec2} we provide a complete proof of Theorem \ref{thm:main2}. In Section \ref{sec3}, we rigorously justify the proof of Theorem \ref{thm:main1} described above, providing the (few) missing details in the above construction. In Section \ref{sec:apps} we present some applications of Theorem \ref{thm:main1}. In Section \ref{sec:revisit}, we revisit Caffarelli's argument and provide an alternative proof of Theorem \ref{thm:main1} for the Brenier map $T_{opt}$ itself. Lastly, in Section \ref{sec:comparing}, we compare between the two maps $T$ and $T_{opt}$, and conclude with some final remarks. \medskip \noindent \textbf{Acknowledgements.} We gratefully acknowledge the support of the Institute for Advanced Study, where this work was initiated, and thank Jean Bourgain and Tom Spencer for their support and interest. We would also like to thank Cedric Villani for his interest and for providing several helpful references during this work, Almut Burchard, Bo'az Klartag and Robert McCann for their interest and illuminating remarks, and Dominic Dotterrer for remarks on terminology. We also thank Haim Brezis, Bob Jerrard, Ki-Ahm Lee, Alessandra Lunardi and Vladimir Maz'ya for their patient help with Proposition \ref{prop:regularity}, and Bernd Kawohl for additional references. Final thanks go out to the anonymous referees, for helpful suggestions which have improved the presentation of this work. \section{Proof of Theorem \ref{thm:main2}} \label{sec2} This section is dedicated to the proof of Theorem \ref{thm:main2}, from which Theorem \ref{thm:main1} easily follows, as explained in the Introduction, and rigorously verified in Section \ref{sec3}. We begin by setting up the notation throughout the paper. Our basic reference is \cite{LadyParabolicBook}, even though our notation varies slightly from the notation used there. We will use $D$ and $\nabla$ interchangeably to denote the derivative operator in $\Real^n$. Given an non-negative integer $k$, we denote by $C^{k}(\Sigma)$ the space of real-valued functions on $\Sigma \subset \Real^n$ with continuous derivatives $D^a f$, for every multi-index $a$ of order $|a| \leq k$, equipped with the usual maximum norm: \[ \norm{f}_{C^{k}(\Sigma)} := \sum_{|a| \leq k} \sup_{x \in \Sigma} |D^{a} f(x)| ~. \] Similarly, the space $C^{k+\alpha}(\Sigma) = C^{k,\alpha}(\Sigma)$ denotes the subspace of functions whose $k$-th order derivatives are H\"{o}lder continuous of order $\alpha \in (0,1]$, equipped with the norm: \[ \norm{f}_{C^{k,\alpha}(\Sigma)} := \norm{f}_{C^{k}(\Sigma)} + \sum_{|a| = k} \sup_{x \neq y \in \Sigma} \frac{|D^a f(x) - D^a f(y)|}{|x-y|^{\alpha}} ~. \] We will say that a continuous function is H\"{o}lder continuous of order 0, in which case $C^{k}(\Sigma)$ indeed coincides with $C^{k,0}(\Sigma)$. When $\Sigma = \Omega \times \Theta$ is a product domain consisting of space $x \in \Omega$ and time $t \in \Theta$ components, we will denote by $C^{k \times l}(\Omega \times \Theta)$ the space of real-valued functions $f$ with continuous (in $\Sigma$) space derivatives $D_x^a$ of order $|a| \leq k$ and time derivatives $D_t^s$ of order $s \leq l$, equipped with the norm: \[ \norm{f}_{C^{k \times l}(\Omega \times \Theta)} := \sum_{{|a| \leq k}} \sup_{z \in \Sigma} |D^{a}_x f(z)| + \sum_{s=0}^{l} \sup_{z \in \Sigma} |D^{s}_t f(z)| ~. \] We will also denote by $C^{(\beta ; \beta/2)}(\Omega \times \Theta)$ the space of real-valued functions $f$ on $\Sigma$ such that for every integer $r,s \geq 0$ with $r + 2s \leq \beta$ and $|a|=r$, $D_x^{a} D_t^{s} f$ is H\"{o}lder continuous in $x$ of order $\min(\beta-(r+2s),1)$ and in $t$ of order $\min(\beta/2-(r/2 +s),1)$. The natural norm on this space is given by: \begin{eqnarray*} \norm{f}_{C^{(\beta ; \beta/2)}(\Omega \times \Theta)} & := & \sum_{r + 2s \leq \lfloor \beta \rfloor} \sum_{|a|=r} \sup_{z \in \Sigma} |D_x^{a} D_t^{s} f(z)| + \\ & + & \sum_{r + 2s = \lfloor \beta \rfloor} \sum_{|a|=r} \sup_{x_1 \neq x_2 \in \Omega , t \in \Theta} \frac{|D_x^{a} D_t^{s} f(x_1,t) - D_x^{a} D_t^{s} f(x_2,t)|}{|x_1 - x_2|^{\beta - (r+2s)}} \\ & + & \sum_{ \lfloor \beta \rfloor -1 \leq r + 2s \leq \lfloor \beta \rfloor} \sum_{|a|=r} \sup_{x \in \Omega , t_1 \neq t_2 \in \Theta} \frac{|D_x^{a} D_t^{s} f(x,t_1) - D_x^{a} D_t^{s} f(x,t_2)|}{|t_1 - t_2|^{\beta/2 - (r/2+s)}} ~. \end{eqnarray*} Lastly, we will denote by $W_p^{(2l;l)}(\Omega \times \Theta)$ for $p \in [1,\infty]$ and $l$ a non-negative integer, the space of functions $f$ on $\Omega \times \Theta$ so that for any integer $r,s \geq 0$ with $r + 2s \leq l$ and $|a|=r$, the distributional derivatives $D_x^{a} D_t^{s} f$ are in $L_p(\Omega \times \Theta)$ (this space is equipped with its usual Sobolev norm, which we will not require explicitly). Finally, we let $F_{loc}(\Sigma)$ denote the space of functions belonging to $F(\Pi)$ for all compact subsets $\Pi$ of $\Sigma$. \subsection{Reduction to smooth $V$} \label{subsec:smooth} Let us start by summarizing several well-known properties of the semi-group $\set{P_t^U}_{t \geq 0}$. From the classical theory of parabolic equations, it follows that for each $t \geq 0$, $P_t^U$ acts linearly on the space $\B(\Real^n)$ of smooth bounded functions on $\Real^n$ to itself (indeed, there exists a unique solution of (\ref{eq:semi-group}) in the class of bounded functions), and hence is a semi-group $P^U_t \circ P^U_s = P^U_{t+s}$. Moreover, by the maximum principle, it follows that $\norm{P_t^U(f)}_{L_\infty} \leq \norm{f}_{L_\infty}$ and that $P_t^U(f) \geq 0$ for any $f \geq 0$ in $\B(\Real^n)$. Since $\int P_t^U(f) d\mu = \int f d\mu$, as easily checked by differentiating in $t$ and using (\ref{eq:semi-group}), it follows by interpolation that $\norm{P_t^U(f)}_{L_p(\mu)} \leq \norm{f}_{L_p(\mu)}$ for all $p \in [1,\infty]$. Consequently, the action of $P^U_t$ extends to all of the $L_p(\mu)$ spaces, clarifying the statement of Theorem \ref{thm:main2}. It follows immediately that it is enough to prove Theorem \ref{thm:main2} for smooth functions $V$. Indeed, any convex function $V : \Real^n \rightarrow \Real \cup \set{+\infty}$ may be pointwise approximated from below by a non-decreasing sequence of smooth convex functions $V_m : \Real^n \rightarrow \Real$, which may be chosen to preserve any symmetry properties satisfied by $V$. In particular, $\exp(-V_m)$ tends to $\exp(-V)$ in $L_1(\mu)$, and so $V_m + c_m$ satisfy the assumptions of Theorem \ref{thm:main2}, where $c_m \rightarrow 0$ denote normalization constants ensuring that $\exp(-(V_m + c_m)) \mu$ are probability measures. Consequently $P_t^U(\exp(-V_m))$ tends to $P_t^U(\exp(-V))$ in $L_1(\mu)$, and since the sequence $P_t^U(\exp(-V_m))$ is pointwise non-increasing (using the positivity of $P_t^U$), it follows that there exists a pointwise limit which coincides with $P_t^U(\exp(-V))$ in $L_1(\mu)$. By assuming that Theorem \ref{thm:main2} holds for smooth functions, it follows that $P_t^U(\exp(-V_m))$ are log-concave: \[ P_t^U(\exp(-V_m))\brac{\frac{x+y}{2}} \geq P_t^U(\exp(-V_m))(x)^{\frac{1}{2}} P_t^U(\exp(-V_m))(y)^{\frac{1}{2}} \;\;\; \forall x,y \in \Real^n ~, \] and this is clearly preserved under pointwise limit. The reduction to the case that $V$ is smooth is complete. \subsection{Reduction to vanishing Dirichlet boundary conditions} Let $B(R)$ denote the open Euclidean ball in $\Real^n$ of radius $R$ centered at the origin, and let $\chi : [0,1] \rightarrow [0,1]$ denote a smooth log-concave (non-increasing) function so that $\chi|_{[0,1)} > 0$, $\chi|_{[0,1/2]} \equiv 1$ and $\chi(1)=0$. \begin{prop} \label{prop:Dirichlet} Let $U \in C^{1,\alpha}_{loc}(\Real^n)$, $V\in C^{2,\alpha}_{loc}(\Real^n)$ and $\exp(-V) \in C^0(\Real^n)$. Assume that for any $R,T > 0$, the solution $f_R(x,t)$ to the parabolic equation: \[ \frac{d}{dt} f_R = \Delta f_R - \scalar{\nabla f_R,\nabla U} ~,~ f_R(x,0) = \exp(-V(x)) \chi(|x|/R) ~ , ~ (x,t) \in B(R) \times [0,T] ~, \] with \emph{vanishing Dirichlet boundary conditions}: \[ f|_{\partial B_R \times [0,T]} \equiv 0 ~, \] is spatially log-concave on $B(R)$ for any $t \in [0,T]$. Then the (unique) bounded solution $f(x,t)$ to the Cauchy problem: \begin{equation} \label{eq:Cauchy} \frac{d}{dt} f = \Delta f - \scalar{\nabla f,\nabla U} ~,~ f(x,0) = \exp(-V(x)) ~, ~ (x,t) \in \Real^n \times [0,\infty) ~, \end{equation} is also spatially log-concave on $\Real^n$ for any $t \geq 0$. \end{prop} \begin{proof} This follows from a standard argument, which we include for completeness. Fix $T > 0$; we will show that $f(x,t)$ is log-concave on $\Real^n$ for any $t \in [0,T]$. By the classical theory of parabolic PDEs (e.g. \cite[Chapter IV, Theorem 10.1]{LadyParabolicBook}), for any $0 < r < r' < R$, we have the following (spatial) interior Schauder-type estimate: \[ \norm{f_R}_{C^{(2+\alpha ; 1+\alpha/2)}(B(r)\times[0,T])} \leq C_1 \norm{f_R(\cdot,0)}_{C^{2+\alpha}(B(r'))} + C_2 \norm{f_R}_{C^0(B(r') \times [0,T])} ~, \] where the constants $C_1,C_2>0$ above depend only on $n,T,\norm{\nabla U}_{C^{0,\alpha}(B(r'))},r,r',\alpha$. By the maximum principle, $\norm{f_R}_{C^0(B(r') \times [0,T])} \leq \norm{\exp(-V)}_{C^0(\Real^n)} < \infty$. And if we assume that $R \geq 1$, since $\chi$ is smooth it follows that $\norm{f_R(\cdot,0)}_{C^{2,\alpha}(B(r'))} \leq C_3 \norm{\exp(-V)}_{C^{2,\alpha}(B(r'))} < \infty$ for some constant $C_3>0$. We conclude that: \[ \forall r > 0, \;\;\; \exists \, C_r > 0 \;\text{ such that } \; \forall R \geq r + 1, \;\;\; \norm{f_R}_{C^{(2 + \alpha; 1+\alpha/2)}(B(r)\times[0,T])} < C_r ~. \] It follows by Arzel\`a--Ascoli compactness that given $r > 0$, we may extract a sequence of $R_m \geq r + 1$ increasing to infinity, so that $f_{R_m}$ converges in $C^{2\times1}(B(r)\times[0,T])$. Applying a standard diagonalization argument, we conclude that there exists a sequence $\set{R_{k}}$ increasing to infinity, so that $f_{R_k}$ converges in $C^{2\times 1}_{loc}(\Real^n\times[0,T])$ to some $f_\infty \in C^{(2+\alpha; 1+\alpha/2)}_{loc}(\Real^n \times [0,T])$ (which is in addition clearly bounded). It follows that $f_\infty$ satisfies (\ref{eq:Cauchy}) on $\Real^n \times [0,T]$, so by the well-known uniqueness of this equation in the class of bounded functions, we deduce that $f_\infty \equiv f$ on $\Real^n \times [0,T]$. But $f_\infty(\cdot,t)$ is clearly log-concave for any $t \in [0,T]$, just from being the pointwise limit of the log-concave functions $f_{R_k}(\cdot,t)$. This concludes the proof. \end{proof} Let $V \in C^{4,\alpha}_{loc}(\Real^n)$ satisfy the assumptions of Theorem \ref{thm:main2}. If we define $V_R\in C^{4,\alpha}(B(R))$ by setting $\exp(-V_R) = \exp(-V(x)) \chi(|x|/R)$ on $B(R)$, we note that the symmetry assumptions of Theorem \ref{thm:main2} remain in tact for $V_R$ on $B(R)$. By Subsection \ref{subsec:smooth} and Proposition \ref{prop:Dirichlet}, Theorem \ref{thm:main2} consequently reduces to the following: \begin{thm} \label{thm:max} Let $U$ be as in Theorem \ref{thm:main1} and let $f_0 \in C^{4,\alpha}(\overline{B(R)})$ denote a positive function on $B(R)$ vanishing on $\partial B(R)$. Assume that on $B(R)$, $f_0 = \exp(-V_0)$, with $V_0$ convex and satisfying our symmetry assumptions (\ref{eq:symmetry}). Then for every $T>0$, the unique solution $f$ to the following parabolic equation on $B(R) \times [0,T]$: \begin{equation} \label{eq:Dirichlet-PDE} \frac{d}{dt} f = \Delta f - \scalar{\nabla f,\nabla U} ~,~ f|_{t=0} = f_0 ~ , ~ f|_{\partial B_R \times [0,T]} \equiv 0 ~, \end{equation} is spatially log-concave, i.e. $f = \exp(-V)$ with $V(\cdot,t)$ convex on $B(R)$ for every $t \in [0,T]$. \end{thm} This reduction step is similar to the one in \cite{DemangePhdThesis}, referenced to us by Cedric Villani, whom we would like to thank. \subsection{Log-Concavity away from the boundary} We proceed to provide a proof of Theorem \ref{thm:max}, modulo some very delicate details which are postponed to the next subsection. As in many previous works on concavity/convexity properties of solutions to elliptic and parabolic PDEs (\cite{LewisConvexRings,KorevaarClassicalPaper,Kennington2,CaffarelliSpruckConvexityProperties,KawohlBook,DiazKawohl,LeeVazquezParabolicNonlinearPDE}), our approach is based on the maximum principle for the second derivative (or its finite difference analogue); other approaches may be found e.g. in \cite{BrascampLiebPLandLambda1,Borell1982QuasiconcavityOfBrownianMotionHittingTime, JapaneseConvexityPreservingSingularParabolicPDE, AlvarezLasryLions,ColesantiSalaniQuasiconcaveEnvelope, BianGuanConstantRank} and the references therein, or in the classical book by B. Kawohl \cite{KawohlBook}. We clarify some of the difficulties which arise in showing log-concavity in the parabolic case and which were omitted in some of these previous works. Another challenge we encounter, is that the condition our parabolic equation must satisfy, so that we can deduce the log-concavity of the solution, in fact assumes that the solution is already log-concave. Hence, arguing in the contrapositive, we must perform our analysis at precisely the \emph{first} time when things go wrong, which again requires some delicate justification. To this end, we avoid using the usual convexity function, introduced by Korevaar \cite{KorevaarClassicalPaper} and employed by many others (see the previously mentioned references or \cite{KawohlBook,LeeVazquezParabolicNonlinearPDE,GrecoKawohl} and the references therein), and work directly with the second derivatives. \begin{proof}[Proof of Theorem \ref{thm:max}] By approximating $f_0$ appropriately and arguing as in Subsection \ref{subsec:smooth}, we may assume that: \begin{equation} \label{eq:slope} \min_{x \in \partial B(R)} |\nabla f_0|(x) > 0 ~; \end{equation} the only difference is that now, due to the boundary conditions, $\norm{f(\cdot,t)}_{L_1(\mu|_{B(R)})}$ will not be preserved, but rather decrease, with time. See also \cite[Lemma 6.1]{GrecoKawohl}, where a similar preliminary step was employed. Fix $T>0$. Since $f_0 \in C^{4,\alpha}(\overline{B(R)})$ and in addition every component of $\nabla U$ is in $C^{2,\alpha}(\overline{B(R)})$, it follows from the classical Schauder theory of parabolic PDEs (e.g. \cite[Chapter IV,Theorem 10.1]{LadyParabolicBook}) that $f \in C_{loc}^{(4+\alpha;2+\alpha/2)}(B(R) \times [0,T])$ (i.e. $f \in C^{(4+\alpha;2+\alpha/2)}(K \times [0,T])$ for every compact subset $K \subset B(R)$), and also that $f \in C^{(4+\alpha;2+\alpha/2)}(\overline{B(R)} \times [\eps,T])$, for any $0<\eps < T$. A crucial point to note is that the latter smoothness of the solution does not extend all the way to the entire boundary $\partial B(R) \times [0,T]$, since our assumption (\ref{eq:slope}) contradicts (in general) the compatibility which is usually assumed between the spatial derivatives of $f_0$ and the time derivatives of our Dirichlet conditions (see Subsection \ref{subsec:boundary}). This difficulty seems unavoidable using this approach, and addressing it requires careful justification of subsequent steps, something which has been omitted in previous works. It also follows from the strong maximum principle (and our initial conditions) that $f>0$ on $B(R) \times [0,T]$, and hence $V \in C^{(4+\alpha;2+\alpha/2)}_{loc}(B(R) \times [0,T])$. One immediately checks that $V$ satisfies the following non-linear parabolic PDE on $B(R) \times [0,T]$: \[ \frac{d}{dt} V = \Delta V - \scalar{\grad V, \grad U} - \scalar{\grad V, \grad V} ~. \] Let $\eps > 0$ and define $\hat{V}\in C^{(4+\alpha;2+\alpha/2)}_{loc}(B(R) \times [0,T])$ as: \[ \hat{V}(x,t) := V(x,t) + \eps \beta(t) \frac{|x|^2}{2} ~, \] where $\beta : [0,T] \rightarrow \Real_+$ denotes a suitable strictly positive smooth function to be determined later on. We claim that for all small enough $\eps > 0$, $\hat{V}(\cdot,t)$ must remain strictly convex for all $t \in [0,T]$, and taking the limit as $\eps \rightarrow 0$, we will conclude that $V(\cdot,t)$ is itself convex, as required. Assume in the contrapositive that this is not so. Let $t_0 \in [0,T]$ denote the infimum over all times $t$ when $\hat{V}(\cdot,t)$ is not strictly convex, so that there exists a sequence $(x_m,t_m,e_m) \in B(R) \times (0,T] \times S^{n-1}$ converging to $(x_0 , t_0 ,e) \in \overline{B(R)} \times [0,T] \times S^{n-1}$ and satisfying $D^2_{e_m,e_m} \hat{V}(x_m,t_m) \leq 0$ (here $S^{n-1}$ denotes the unit sphere in $\Real^n$, identified with the unit sphere in the tangent spaces $T_{x_m}\Real^n$). The most delicate part of the proof will be presented in Proposition \ref{prop:boundary} in the next subsection, where it will be shown that some further regularity estimates of $f$ up to the boundary, together with (\ref{eq:slope}) and the strict convexity of $\partial B(R)$, imply that necessarily $x_0 \notin \partial B(R)$. It follows by continuity of the second derivative of $\hat{V}$ in $B(R) \times [0,T]$ and the minimality of $t_0$ that $D^2_{e,e} \hat{V}(x_0,t_0) = 0$, and therefore $t_0 > 0$ (since at time $t=0$, $\hat{V}(\cdot,t)$ is clearly strictly convex). Moreover, $x_0 \in B(R)$ is a local minimum point, and hence: \begin{equation} \label{eq:max1} D D^2_{e,e} \hat{V} (x_0,t_0) = 0 ~, ~ \Delta D^2_{e,e} \hat{V} (x_0,t_0) \geq 0 ~,~ \frac{d}{dt} D^2_{e,e}\hat{V} (x_0,t_0) \leq 0 ~, \end{equation} where $D$ denotes the space derivative. Since $0$ is the minimum value for the function $e \rightarrow D^2_{e,e} \hat{V}(x_0,t_0)$, it follows that it must be an eigenvalue of $D^2 \hat{V}(x_0,t_0)$, and that $e$ is a corresponding eigenvector: \begin{equation} \label{eq:max2} D D_e \hat{V}(x_0,t_0) = D^2 \hat{V}(x_0,t_0) e = 0 ~ , ~ \text{and hence } ~ D D_e V(x_0,t_0) = - \eps \beta(t_0) e ~. \end{equation} Using (\ref{eq:max1}), we must have at $(x_0,t_0)$: \begin{equation} \label{eq:max3} (d/dt - \Delta)(D^2_{e,e} \hat{V}) \leq 0 ~. \end{equation} We will show that under our assumptions on $U$ and the definition of $t_0$, the latter value must be strictly positive, obtaining the desired contradiction and concluding the proof. Indeed, at a general point $(x,t)$: \begin{eqnarray*} & & (d/dt - \Delta)(D^2_{e,e} \hat{V}) = D^2_{e,e}((d/dt - \Delta)(\hat{V})) \\ &=& D^2_{e,e}(\eps \beta'(t) |x|^2/2 - n \eps \beta(t) - \scalar{\grad V,\grad U} - \scalar{\grad V,\grad V}) = \eps \beta'(t) - \sscalar{D D^2_{e,e} V , DU} \\ & & - 2 \scalar{D D_e V, D D_e U} - \scalar{DV , D D^2_{e,e} U} - 2 \scalar{D D^2_{e,e} V , DV} - 2 \scalar{D D_e V, D D_e V} ~. \end{eqnarray*} At $(x_0,t_0)$, using $(\ref{eq:max1})$ and $(\ref{eq:max2})$, we see that: \begin{eqnarray*} (d/dt - \Delta)(D^2_{e,e} \hat{V})(x_0,t_0) = \eps \beta'(t_0) + 2 \eps \beta(t_0) D^2_{e,e} U - 2 \eps^2 \beta(t_0)^2 - \scalar{DV , D D^2_{e,e} U} \\ = \eps \beta'(t_0) + 2 \eps \beta(t_0) D^2_{e,e} U - 2 \eps^2 \beta(t_0)^2 + \eps \beta(t_0) \scalar{x, D D^2_{e,e} U} - \sscalar{D\hat{V} , D D^2_{e,e} U} \\ \geq \eps \beta'(t_0) - (2 \eps \beta(t_0) M_2 + 2 \eps^2 \beta(t_0)^2 + \eps \beta(t_0) R M_3) - D^3 U(e,e,D \hat{V}) ~, \end{eqnarray*} where $M_2 := \sup_{x \in B(R), \xi \in S^{n-1}} D^2_{\xi,\xi} U(x)$ and $M_3 := \sup_{x \in B(R), \xi \in S^{n-1}} |(D^3 U)|_{x}(\xi,\xi,\frac{x}{|x|})|$. Note that by the definitions of $t_0$ and $x_0$, $D^2_{\xi,\xi}\hat{V}(x,t_0) \geq D^2_{e,e}\hat{V}(x_0,t_0) = 0$, so $\hat{V}(\cdot,t)$ is still convex on $B(R)$ at time $t = t_0$. Also note that since $U$, $f_0$ (and $B(R)$) are all invariant under the action of $O(E_1,\ldots,E_k)$, and since the Laplace operator commutes with the entire orthogonal group, it follows easily that $f \circ G$ is also a solution to (\ref{eq:Dirichlet-PDE}) for any $G \in O(E_1,\ldots,E_k)$. The uniqueness of the solution implies that $f(\cdot,t)$ (and hence $V(\cdot,t)$ and $\hat{V}(\cdot,t)$) are also invariant under the action of this subgroup, and hence satisfy our symmetry assumptions for all $t \geq 0$. We will see in Proposition \ref{prop:geom} below that for any convex function $F : \Real^n \rightarrow \Real$ satisfying our symmetry assumptions, the condition on $U$ implies that $(D^3 U)|_{x}(\xi,\xi,D F(x)) \leq 0$ for any $x \in \Real^n$ and $\xi \in S^{n-1}$. Therefore, in order to arrive to a contradiction with (\ref{eq:max3}), it is enough to show that for small enough $\eps>0$ and an appropriate choice of $\beta$, we have: \[ \beta'(t_0) - (2 \beta(t_0) M_2 + 2 \eps \beta(t_0)^2 + \beta(t_0) R M_3) > 0 ~. \] Indeed, this is satisfied on $[0,T]$ by setting $\beta(t) := \exp((2M_2 + R M_3 + 1) t)$ and letting $\eps < 1/(2\beta(T))$. This completes the contradiction and concludes the proof, modulo Propositions \ref{prop:geom} and \ref{prop:boundary} below. \end{proof} We conclude this subsection with the proof of the following proposition, which is the only place where we use our structural assumptions on $U$ and $V$. In fact, the assumption that $U$ is convex may be omitted in all instances below (see Section \ref{sec:comparing} for more on this). \begin{prop} \label{prop:geom} If $U$ and $V$ satisfy the assumptions of Theorem \ref{thm:main1} then: \[ (D^3 U)|_{x}(\xi,\xi,\grad V(x)) \leq 0 \;\;\; \forall x \in \Real^n ~ \forall \xi \in S^{n-1} ~. \] \end{prop} The proposition follows immediately from the following two lemmata, which we formulate separately for later use: \begin{lem} \label{lem:geom1} Let $U$ satisfy the assumptions of Theorem \ref{thm:main1}. Then $(D^3 U)|_{x}(\xi,\xi,\theta) \leq 0$, for any $x \in \Real^n$, $\xi \in S^{n-1}$ and $\theta \in S^{n-1}$ such that: \begin{equation} \label{eq:theta-cond} \forall i=1,\ldots,k \;\;\; \exists a_i \geq 0 \;\;\; \text{so that} \;\;\; Proj_{E_i} \theta = a_i Proj_{E_i} x ~. \end{equation} \end{lem} \begin{lem} \label{lem:geom2} Let $V$ satisfy the assumptions of Theorem \ref{thm:main1}. Then for any $x \in \Real^n$, $\theta = \nabla V(x)$ satisfies (\ref{eq:theta-cond}). \end{lem} \begin{proof}[Proof of Lemma \ref{lem:geom1}] Let $\varrho_i : E_i \rightarrow \Real$ be given by $\varrho_i(x) = \rho_i(|x|)$, $i=1,\ldots,k$. Taking the third derivative of $U$, the quadratic term in (\ref{eq:U}) disappears and we are left with: \[ \left . (D^3 U) \right |_{x}(\xi,\xi,\theta) = \sum_{i=1}^k \left . (D^3_{E_i} \varrho_i) \right |_{Proj_{E_i} x}(Proj_{E_i} \xi,Proj_{E_i} \xi,Proj_{E_i} \theta) ~. \] Let us show that each summand is non-positive. Denote: \[ x_i := Proj_{E_i} x ~,~ \xi_i := Proj_{E_i} \xi ~,~ \xi_i^r := Proj_{x_i} \xi_i ~,~ \xi_i^t := Proj_{x_i^\perp} \xi_i ~,~ \theta_i := Proj_{E_i} \theta ~. \] If $x_i = 0$ then $\theta_i = 0$ and hence the $i$-th summand is also $0$, so we may assume that $x_i \neq 0$. Using (\ref{eq:theta-cond}), an elementary calculation yields: \[ \left . (D^3_{E_i} \varrho_i) \right |_{x_i} (\xi_i,\xi_i,\theta_i) = \brac{ \rho_i'''(|x_i|) |\xi_i^r|^2 + \brac{\rho_i''(|x_i|) - \frac{\rho_i'(|x_i|)}{|x_i|}} \frac{|\xi_i^t|^2}{|x_i|} } a_i |x_i| ~. \] Since $t \mapsto \rho_i(|t|)$ is a $C^3$ function, we see that $\rho_i'(0) = 0$. Since $\rho'''_i \leq 0$ on $\Real_+$, meaning that $\rho'_i$ is concave there, we deduce that also $\rho_i''(t) \leq (\rho_i'(t)-\rho_i'(0))/t = \rho_i'(t)/t$ for all $t > 0$. This implies that the term in brackets on the rand-hand side above is non-positive, and since $a_i \geq 0$, the entire expression is non-positive as well, as claimed. \end{proof} \begin{proof}[Proof of Lemma \ref{lem:geom2}] Denote as usual $x_i = Proj_{E_i} x$, $i=0,1,\ldots,k$. Let us verify (\ref{eq:theta-cond}) for each $i=1,\ldots,k$. It is easy to see that the symmetries of $V$ ensure that $D_i V(x) := Proj_{E_i} \nabla V(x)$ lies in the one-dimensional subspace spanned by $x_i$. Hence if $x_i = 0$, then $D_i V(x) =0$ and (\ref{eq:theta-cond}) is satisfied trivially for that $i$, so we may assume otherwise. Denoting: \[ D_i V(x) =: D_i^r V(x) \frac{x_i}{|x_i|} ~, \] it remains to verify that $D_i^r V(x) \geq 0$ when $x_i \neq 0$. The symmetries of $V$ together with its convexity together imply that the following (convex) slice of $V$'s sub-level set at $x$: \[ A(x):=\set{z \in E_0^\perp ; V(x_0 + z) \leq V(x)} ~, \] contains the product set $B_{E_1}(|x_1|) \times \ldots \times B_{E_k}(|x_k|)$, where $B_{E_i}(r)$ denotes the Euclidean ball of radius $r$ in $E_i$. Geometrically, this means that the latter product set lies entirely on one side of the tangent plane to $A(x)$ at $Proj_{E_0^\perp} x$, or more precisely, that: \[ \scalar{Proj_{E_0^\perp} \grad V(x) , R(x) - x} \leq 0 \;\;\; \forall R \in O(E_1,\ldots,E_k) ~. \] Recalling that $Proj_{E_0^\perp} \grad V(x) = \sum_{i=1}^k D_i^r V(x) x_i$ and choosing $R_i \in O(E_1,\ldots,E_k)$ to be the reflection in $E_i$, defined by $R_i(x) = x - 2 x_i$, we conclude that: \[ D_i^r V(x) |x_i|^2 \geq 0 \;\;\; \forall i=1,\ldots,k ~. \] Since we assumed that $x_i \neq 0$, it follows that $D_i^r V(x) \geq 0$, as required. \end{proof} \subsection{Log-Concavity near the boundary} \label{subsec:boundary} To complete the proof of Theorem \ref{thm:max}, we must show that $x_0 \notin \partial B(R)$. Recalling the definition of $x_0$, this clearly follows from: \begin{prop}\label{prop:boundary} $D^2 V(x,t) \geq 0$ in a neighborhood of $\partial B(R) \times [0,T]$. \end{prop} The proof of Proposition \ref{prop:boundary} will be given at the end of this section, but first we explain the subtle regularity issue one is required to address here. Recall that the classical theory guarantees that under the assumptions of Theorem \ref{thm:max}, $f \in C_{loc}^{(4+\alpha;2+\alpha/2)}(B(R) \times [0,T])$ (i.e. $f \in C^{(4+\alpha;2+\alpha/2)}(K \times [0,T])$ for every compact subset $K \subset B(R)$), and also that $f \in C^{(4+\alpha;2+\alpha/2)}(\overline{B(R)} \times [\eps,T])$, for any $0<\eps < T$. However, the latter smoothness does not extend all the way to the ``corner'' $\partial B(R) \times \set{0}$, since in general we cannot guarantee the necessary and sufficient compatibility conditions: \begin{equation} \label{eq:compat} L^i(f_0)|_{\partial B(R)} \equiv 0 \;\;\;\;\; i=1,2 \end{equation} (here $L^i$ denotes the iterated application of the operator $L$). This prevents a straightforward application of standard arguments for deducing Proposition \ref{prop:boundary}, and so we consequently need to obtain some delicate regularity estimates \emph{up to the boundary} for the solution $f$ to (\ref{eq:Dirichlet-PDE}), which are given in Proposition~\ref{prop:regularity} below. \medskip To outline the proof and properly motivate Proposition~\ref{prop:regularity}, observe that: \[ D^2 V = -D^2 \log f = \frac{ - f D^2 f + \nabla f \otimes \nabla f}{f^2 } ~. \] Using Hopf's maximum principle and continuity of $\nabla f$ (see Proposition~\ref{prop:regularity} (1)) we see below that $\nabla f$ is bounded uniformly away from zero near $\partial B(R) \times [0, T]$. Therefore, the term $\nabla f \otimes \nabla f$ is uniformly positive definite when restricted to the normal direction (relative to $\partial B(R)$). In addition, the gradient bound implies that $f$ decays linearly to $0$ near the boundary, and one can show that $-fD^2 f$ decays uniformly to zero near $\partial B(R) \times [0,T]$ (see Proposition~\ref{prop:regularity} (2)). It follows that $D^2 V$ restricted to the normal direction is uniformly positive definite near $\partial B(R) \times [0,T]$. On the other hand, since $\partial B(R)$ is the zero level set of $f$, the uniform convexity of $\partial B(R)$ and the uniform lower bound on $|\nabla f|$ together imply that $-D^2 f$ restricted to the tangential directions is uniformly positive definite along $\partial B(R) \times [0, T]$. Since the tangential second derivatives of $f$ are uniformly continuous (see Proposition~\ref{prop:regularity} (3)), it follows that $D^2 V$ restricted to the tangential directions is uniformly positive definite in a neighborhood of $\partial B(R) \times [0,T]$. Mixed derivatives are controlled similarly. We now proceed with providing the precise details. We begin with: \begin{prop} \label{prop:regularity} Under the assumptions of Theorem \ref{thm:max}: \begin{enumerate} \item $f \in C^{(1+\beta; (1+\beta)/2)}(\overline{B(R)} \times [0,T])$ for all $\beta \in (0,1)$. \item For any $\eps > 0$ there exists a $C_\eps > 0$ so that for any $\lambda \in (0,R)$: \begin{equation} \label{eq:weak-C^2} \sup_{t \in [0,T]} \norm{f(\cdot,t)}_{C^{2}(\overline{B(R-\lambda)})} \leq \frac{C_\eps}{\lambda^{\eps}} ~. \end{equation} \item If $n \geq 2$, the spatial derivatives of $f$ in the non-radial directions are $C^{1,\delta}(\overline{B(R)})$ uniformly in $t \in [0,T]$, for any $\delta \in (0,1)$. In other words, for any $\delta \in (0,1)$ there exists a finite constant $C_\delta>0$, so that for any smooth unit vector field $\xi$ on $\overline{B(R)}$ such that $\scalar{\xi(x),x}\equiv 0$: \[ \sup_{t \in [0,T]} \norm{\scalar{\grad f(\cdot,t),\xi(\cdot)}}_{C^{1,\delta}(\overline{B(R)})} \leq C_\delta ~; \] (in fact, we actually have $\norm{\scalar{\grad f,\xi}}_{C^{(1+\delta;(1+\delta)/2)}(\overline{B(R)}\times[0,T])} \leq C_\delta$ ). \end{enumerate} \end{prop} \begin{rem} We were informed by Ki-Ahm Lee and Vladimir Maz'ya that it should actually be true that: \[ \sup_{t \in [0,T]} \norm{f(\cdot,t)}_{C^{1,1}(\overline{B(R)})} < \infty ~, \] but we will not insist on this here since the easier weaker estimate (\ref{eq:weak-C^2}) suffices for our purposes. \end{rem} \begin{proof}[Proof of Proposition \ref{prop:regularity}] 1. The first assertion follows from standard regularity theory. Even if the compatibility conditions (\ref{eq:compat}) do not necessarily hold, it follows by \cite[Theorem 5.1.11 (ii)]{LunardiBook} that when $f_0 \in C^{1,\beta}(\overline{B(R)})$ for some $\beta \in (0,1)$ and $f_0|_{\partial B(R)} \equiv 0$, then: \begin{equation} \label{eq:reg-1} f \in C^{(1+\beta;(1+\beta)/2)}(\overline{B(R)} \times [0,T]) ~. \end{equation} Alternatively, one may employ the Sobolev regularity theory for parabolic PDEs (e.g. \cite[Chapter IV, Theorem 9.1 and subsequent Corollary]{LadyParabolicBook}), which ensures that $f \in W_p^{(2;1)}(\overline{B(R)} \times [0,T])$ for all $p \in (1,\infty)$. Consequently, (\ref{eq:reg-1}) follows by a variant of Morrey's embedding theorem (e.g. \cite[Chapter II, Lemma 3.3]{LadyParabolicBook}). \bigskip \noindent 2. This may be deduced from \cite[Theorem 5.15]{LiebermanBook} by considering weighted H\"{o}lder spaces. To avoid these, one may proceed as follows. Applying a standard Schauder-type interior estimate, if $f_0 \in C^{2,\gamma}(\overline{B(R)})$ and each component of $\grad U$ is in $C^{0,\gamma}(\overline{B(R)})$, one checks (see e.g. \cite[p. 355]{LadyParabolicBook}) that: \begin{equation} \label{eq:reg-2} \norm{f}_{C^{(2+\gamma;1+\gamma/2)}(\overline{B(R-\lambda)}\times[0,T])} \leq \frac{C_\gamma}{\lambda^{2+\gamma}} \;\;\; \forall \lambda \in (0,R) ~. \end{equation} Combining (\ref{eq:reg-1}) and (\ref{eq:reg-2}), we deduce under the assumptions of Theorem \ref{thm:max}, that for all $\lambda \in (0,R)$: \begin{eqnarray*} \sup_{t \in [0,T]} \norm{f(\cdot,t)}_{C^{1,\beta}(\overline{B(R-\lambda)})} \leq B_\beta \;\;\; \forall \beta \in (0,1) ~ ; \\ \sup_{t \in [0,T]} \norm{f(\cdot,t)}_{C^{2,\gamma}(\overline{B(R-\lambda)})} \leq \frac{C_\gamma}{\lambda^{2+\gamma}} \;\;\; \forall \gamma \in (0,1) ~. \end{eqnarray*} Since $\partial B(R-\lambda)$ is uniformly smooth for all $\lambda \in (0,R/2)$, one can use interpolation in the spaces of H\"{o}lder differentiable functions (see Lunardi \cite[Corollary 1.2.19,1.2.7]{LunardiBook}), and obtain for any $\eta \in (0,\gamma)$ and $\lambda$ in this range: \[ \sup_{t \in [0,T]} \norm{f(\cdot,t)}_{C^{2,\eta}(\overline{B(R-\lambda)})} \leq A_{2+\gamma,2+\eta,1-\beta} B_{\beta}^{\frac{\gamma-\eta}{\gamma+1-\beta}} C_{\gamma}^{\frac{1-\beta+\eta}{\gamma+1-\beta}} \lambda^{-\frac{(2+\gamma)(1-\beta+\eta)}{\gamma+1-\beta}} ~. \] By modifying the constants above, the bound remains valid for all $\lambda \in (0,R)$. Choosing $\eta > 0$ and $1-\beta > 0$ very small, the second part of Proposition \ref{prop:regularity} follows. \bigskip \noindent 3. This part is obtained by first flattening the boundary $\partial B(R)$ near a point, and then applying the standard parabolic regularity theory to the resulting PDE for $D_\tau f$, where $\tau$ denotes a vector parallel to the flattened boundary. This procedure is standard, and the details are provided for the reader's convenience. Let us fix an orthogonal basis $e_1,\ldots,e_n$ of $(\Real^n,|\cdot|)$ and a direction $\xi_0 \in S^{n-1}$. Let $T : \overline{B(R)} \rightarrow \overline{\Omega}$ denote a smooth diffeomorphism so that $T$ coincides with the usual Cartesian-to-polar change of coordinates on the half-annulus $A_+ := B(R) \setminus \overline{B(R/2)}) \cap \set{x \in \Real^n ; \scalar{x,\xi_0} > 0}$. Now consider the PDE satisfied by $g = f \circ T^{-1}$ on $\Omega$. Since both $T$ and $T^{-1}$ are smooth and in particular Lipschitz, it is easy to check that $g$ satisfies a uniformly parabolic PDE on $\Omega \times [0,T]$ of the form: \begin{equation} \label{eq:polar-PDE} \frac{d}{dt} g = \sum_{i,j} a_{i,j} D^2_{i,j} g + \sum_i b_i D_i g ~, \end{equation} where $a_{i,j} = a_{i,j}(y)$ is a uniformly positive-definite smooth matrix and $b_i = b_i(y)$ have the same smoothness as $\grad U$, i.e. $b_i \in C^{2,\alpha}(\overline{\Omega})$. Moreover, since in polar-coordinates: \[ \Delta = r^{-n+1} \frac{\partial}{\partial r} (r^{n-1} \frac{\partial}{\partial r}) + \frac{1}{r^2} \Delta_{S^{n-1}} ~, \] we see that on $T(A_+)$, if we use the natural basis $y = (\theta_1,\ldots,\theta_{n-1},r)$ to write (\ref{eq:polar-PDE}), we actually have: \begin{equation} \label{eq:polar-matrix} a_{i,j}(\theta_1,\ldots,\theta_{n-1},r) = \begin{cases} \delta_{i,j} & i=n \\ \frac{\delta_{i,j}}{r^2} & i=1,\ldots,n-1 \end{cases} ~. \end{equation} Finally, since $T$ is a diffeomorphism, $T(\partial B(R)) = \partial \Omega$, and hence the boundary conditions are given by: \[ g|_{t=0} = g_0 := f_0 \circ T^{-1} \;\;\;, \;\;\; g|_{\partial \Omega \times [0,T]} \equiv 0 ~. \] The usual regularity theory ensures that $g \in C_{loc}^{(4+\alpha;2+\alpha/2)}(\Omega \times [0,T])$, and as in the first part, it follows that: \begin{equation} \label{eq:g-reg} g \in C^{(1+\delta;(1+\delta)/2)}(\overline{\Omega} \times [0,T]) \;\;\; \forall \delta \in (0,1) ~. \end{equation} Now take the spatial derivative of (\ref{eq:polar-PDE}) in a direction $\tau \in \text{span}(e_1,\ldots,e_{n-1})$. Denoting $g_\tau := D_\tau g$, we obtain that in $\Omega \times [0,T]$: \[ \frac{d}{dt} g_\tau = \sum_{i,j} a_{i,j} D^2_{i,j} g_\tau + \sum_{i,j} D_\tau a_{i,j} D^2_{i,j} g + \sum_i b_i D_i g_\tau + \sum_i D_\tau b_i D_i g ~. \] By (\ref{eq:g-reg}), the fourth term on the right hand side, which we denote by $h$, is in $C(\overline{\Omega}\times[0,T])$ (and in fact better). The second term above contains mixed second derivatives of $g$, but fortunately in $T(A_+)$, the matrix $a_{i,j}(y)$ is given by (\ref{eq:polar-matrix}), and hence $D_\tau a_{i,j}(y) = 0$. We conclude that in $T(A_+) \times [0,T]$, $g_\tau$ satisfies the following uniformly parabolic PDE: \begin{equation} \label{eq:g-tau-PDE} \frac{d}{dt} g_\tau = \sum_{i,j} a_{i,j}(y) D^2_{i,j} g_\tau + \sum_i b_i(y) D_i g_\tau + h(y,t) ~, \end{equation} and that: \[ g_\tau|_{t=0} = D_\tau g_0 \;\;\; , \;\;\; g_\tau|_{(\partial T(A_+) \cap \partial \Omega) \times [0,T]} \equiv 0 ~. \] Employing the standard regularity theory, it follows as in the first part that $g_\tau \in C^{(1+\delta;(1+\delta)/2)}(\overline{\Theta}\times[0,T])$ for any $\delta \in (0,1)$ and open subset $\Theta \subset \Omega$ with smooth boundary, which is in addition bounded away from $\partial T(A_+) \setminus \partial \Omega$. Recalling that $g = f \circ T^{-1}$ and that $T$ is a polar change-of-coordinates on $T(A_+)$, the third assertion of the proposition follows on $(B(R\xi_0,a) \cap \overline{B(R)}) \times [0,T]$ for some small enough $a>0$ (here $B(z,a)$ denotes the ball of radius $a$ centered at $z$). By following the bounds obtained in the proof, one may check that these do not depend on the choice of $\xi_0$ or the non-radial direction $\tau$. By compactness (or using the fact that actually $a>0$ does not depend on $\xi_0$), the assertion follows on a uniform neighborhood of $\partial B(R) \times [0,T]$, and the classical theory takes care of the interior regularity. This completes the proof. \end{proof} \begin{proof}[Proof of Proposition \ref{prop:boundary}] Recall that by the classical theory, $f(\cdot,t) \in C^{4,\alpha}(\overline{B(R)})$ for every $t \in [0,T]$. The second fundamental form of a spatial level set $M$ of $f$ at a point $(x,t)$ with $\nabla f(x,t) \neq 0$, i.e. $M = M_{v,t} := \set{z \in \overline{B(R)} ; f(z,t) = v}$ where $v = f(x,t)$, is given by: \[ II_{M}(x) = - \left . D \frac{\nabla f}{|\nabla f|} \right |_{T_x M} = - \left . \frac{D^2 f}{|\nabla f|} \brac{Id - \frac{\nabla f}{|\nabla f|} \otimes \frac{\nabla f}{|\nabla f|}} \right |_{T_x M} = - \left . \frac{D^2 f}{|\nabla f|} \right |_{T_x M} ~. \] Since we assumed in (\ref{eq:slope}) that $|\nabla f_0 | > 0$ on $\partial B(R)$ and since $\nabla f$ is (uniformly) continuous on $\overline{B(R)}\times[0,T]$ by Proposition \ref{prop:regularity} (1), it follows that there exists some $T_0 > 0$ so that $|\nabla f| \geq c' > 0$ on all of $\partial B(R) \times [0,T_0]$. By the strong maximal principle and Hopf's lemma in the parabolic setting (see e.g. \cite[Chapter 2, Theorem 14]{FriedmanPDEBook}), $|\nabla f| \geq c'' > 0$ on all of $\partial B(R) \times [T_0,T]$, and by the uniform continuity of $\nabla f$ we conclude that there exists $R' \in (0,R)$ and $c,C >0$ so that: \[ 0 < c \leq -\scalar{\nabla f(x,t),\frac{x}{|x|}} \leq |\nabla f(x,t)| \leq C , \;\;\; \forall |x| \in [R',R] \;\;\; \forall t \in [0,T] ~, \] and hence: \begin{equation} \label{eq:good-slope2} c (R - |x|) \leq f(x,t) \leq C (R- |x|), \;\;\; \forall |x| \in [R',R] \;\;\; \forall t \in [0,T] ~. \end{equation} Since the level set $M_{0,t}$ coincides with $\partial B(R)$ for all $t \in [0,T]$ ($f>0$ in $B(R)\times [0,T]$ by the strong maximum principle), it follows that: \[ - \left . \frac{D^2 f}{|\nabla f|} \right |_{x^\perp} = \frac{1}{R} \left . Id \right|_{x^\perp} \;\;\; \forall (x,t) \in \partial B(R) \times [0,T] ~, \] where $x^\perp$ is identified with $T_x \partial B(R)$. By Proposition \ref{prop:regularity} (3), the second spatial derivatives of $f$ involving a non-radial direction are (uniformly) continuous on $\overline{B(R)} \times [0,T]$, and so we deduce that there exists some $R'' \in [R',R)$ so that: \[ -D^2_{\tau,\tau} f(x,t) \geq \frac{c}{2 R} \text{ and } -D^2_{\tau,\frac{x}{|x|}} f(x,t) \geq -B \;\;\;\; \forall |x| \in [R'',R] \;\;\; \forall t \in [0,T] \;\;\; \forall \tau \in S^{n-1} \cap x^\perp ~. \] where: \[ B := \max\set{ \abs{D^2_{\tau,\frac{x}{|x|}} f(x,t)} \; ; \; x \in \overline{B(R)} , \tau \in S^{n-1} \cap x^\perp , t \in [0,T] } < \infty ~. \] Since the tangential derivatives of $f$ vanish on $\partial B(R)$, it also follows that: \[ |\scalar{\grad f(x,t), \tau}| \leq B (R - |x|) \;\;\;\; \forall |x| \in [R'',R] \;\;\; \forall t \in [0,T] \;\;\; \forall \tau \in S^{n-1} \cap x^\perp ~. \] Lastly, fixing $\eps \in (0,1)$, it follows by Proposition \ref{prop:regularity} (2) and (\ref{eq:good-slope2}) that: \[ -f(x,t) D^2_{\xi,\xi} f (x,t) \geq - C C_\eps (R - |x|)^{1-\eps} \;\;\;\; \forall |x| \in [R'',R] \;\;\; \forall \xi \in S^{n-1} \;\;\; \forall t \in [0,T] ~. \] We are now ready to bound $D^2 V$, using: \[ D^2 V = -D^2 \log f = \frac{ - f D^2 f + \nabla f \otimes \nabla f}{f^2 } ~. \] Given $x$ with $|x| \in [R'',R]$ and a direction $\xi \in S^{n-1}$, write $\xi = \cos(\theta) \tau + \sin(\theta) \rho$, where $\rho = x/|x|$ and $\scalar{\tau,\rho} = 0$. For the purpose below, we can assume without loss of generality that $\theta \in [0,\pi/2]$. At the point $(x,t)$, denoting in addition $d = R - |x|$, we have by all the estimates above: \begin{eqnarray*} f^2 D^2_{\xi,\xi} V & = & \cos^2(\theta) (- f D^2_{\tau,\tau} f + \scalar{\grad f,\tau}^2 ) + \sin^2(\theta) (- f D^2_{\rho,\rho} f + \scalar{\grad f,\rho}^2 ) \\ & + & 2 \sin(\theta) \cos(\theta) ( - f D^2_{\tau,\rho} f + \scalar{\grad f,\tau} \scalar{\grad f,\rho}) \\ & \geq & \cos^2(\theta) c d \frac{c}{2R} + \sin^2(\theta) ( - C C_\eps d^{1-\eps} + c^2) + 2 \cos(\theta) \sin(\theta) (-C B d - C B d) ~. \end{eqnarray*} We see that if $d \in [0,d_0]$ for some small enough $d_0 \in (0,R-R'']$, we have for some $p,q,r,p',q' > 0$: \begin{eqnarray*} f^2 D^2_{\xi,\xi} V & \geq & \cos^2(\theta) p d + \sin^2(\theta) q - 2 \cos(\theta)\sin(\theta) r d \\ & \geq & \cos^2(\theta) \frac{p}{2} d + \sin^2(\theta) \brac{q - \frac{2 r^2}{p} d} \geq \cos^2(\theta) p' d + \sin^2(\theta) q' ~, \end{eqnarray*} and so when $d \in (0,d_0]$: \[ D^2_{\xi,\xi} V \geq \frac{\cos^2(\theta) p' d + \sin^2(\theta) q'}{C^2 d^2} > 0 ~; \] (indeed, this behaviour as a function of $\theta,d$ is the best one can expect). We conclude that $D^2_{\xi,\xi} V(x,t) \geq 0$ (and in fact, tends to $+\infty$ uniformly in $d$) for all $|x| \in [R-d_0,R]$, $t \in [0,T]$ and $\xi \in S^{n-1}$. The proof is complete. \end{proof} \section{Tying up loose ends} \label{sec3} In this section, we provide a complete justification of the proof of Theorem \ref{thm:main1}, described in the Introduction. We proceed with the same notations used there. The main technical points which we address in this section are showing that the flow map $S_t$ is globally well-defined on $\mathbb{R}^n$ (see Lemma~\ref{lem:hessian of V bound} and its preceding discussion), that the pushed-forward measure $\nu_t := (S_{t})_* \nu= P_t^U(\exp(-V)) \mu$ converges to $\mu$ (see Lemma~\ref{lem:convergence}), and that the inverse map $T_t = S_t^{-1}$ converges (to a contracting map) as $t \to \infty$ (see Lemma~\ref{lem:approx}) . Let $U,V$ be as in Theorem \ref{thm:main1}. We assume further that $V$ is sufficiently smooth (e.g. $V \in C^{4,\alpha}(\Real^n)$ is more than enough), and that: \begin{equation} \label{eq:sec3-assumptions} \norm{\nabla V}_{C^{1,\alpha}(\Real^n)} < \infty \text{ and } \norm{D^3 U}_{L_\infty} < \infty ~. \end{equation} We will see how to obtain the general case at the very end of this section. First, since $\exp(-V) \in C^{4,\alpha}(\Real^n)$ and $U \in C^{3,\alpha}_{loc}(\Real^n)$, the classical regularity theory of parabolic PDEs (e.g. \cite{LadyParabolicBook}) ensures that $f(x,t):=P_t^U(\exp(-V))(x)$, as the unique (bounded) solution to the Cauchy problem: \begin{equation} \label{eq:sec3-PDE} \frac{d}{dt} f = L f ~,~ f|_{t=0} = \exp(-V) ~, \end{equation} is $C^{(4+\alpha ; 2+\alpha/2)}_{loc}(\Real^n \times [0,\infty))$, and the strong maximum principle ensures that $f(x,t)$ is strictly positive for all $t \in [0,\infty)$. Consequently, the advection field $W_t := -\nabla \log P_t^U(\exp(-V))$ is $C^{(3+\alpha;(3+\alpha)/2)}_{loc}(\Real^n \times [0,\infty))$. In particular, the maps $S_t$ defined by: \begin{equation} \label{eq:ODE} \frac{d}{dt} S_t(x) = W_t(S_t(x)) ~,~ S_0 = Id ~, \end{equation} are indeed \emph{locally} well-defined as a solution to a flow along a locally Lipschitz vector field (e.g. \cite[Proposition 1.56]{GHLBookEdition2}): for any compact subset $C \subset \Real^n$, there exists $t(C) > 0$, so that (\ref{eq:ODE}) has a solution for any $(x,t) \in C \times [0,t(C))$. To ensure that the maps $S_t$ are globally well-defined, it is enough to show that for any $T > 0$, $W_t(x)$ is \emph{globally} spatially Lipschitz for all $t \in [0,T]$, i.e. $|D W_t(x)| < C(T)$ for all $(x,t) \in \Real^n \times [0,T]$: \begin{lem}\label{lem:hessian of V bound} Assuming (\ref{eq:sec3-assumptions}), for all $T>0$, $D^2 \log P_t^U(\exp(-V))(x)$ is uniformly bounded in $\Real^n \times [0,T]$. \end{lem} \begin{proof} We denote by abuse of notation $V = V(x,t) = - \log P_t^U(\exp(-V))(x)$ and $V_t = V(\cdot,t)$. Since $D^2 V \geq 0$ by Theorem \ref{thm:main2}, it suffices to show a uniform bound on $Z = \Delta V$. Recall from Section \ref{sec2} that $V$ satisfies: \begin{equation} \label{eq:V-PDE} \frac{d}{dt} V = \Delta V - \scalar{\nabla V,\nabla V} - \scalar{\nabla V,\nabla U} ~,~ V|_{t=0} = V_0 ~. \end{equation} A direct calculation gives: \begin{align*} \frac{d}{dt} Z = & \Delta Z - 2 \scalar{\nabla Z , \nabla V} - \scalar{\nabla Z,\nabla U}\\ & - 2 tr((D^2 V)^* D^2 V) - 2 tr((D^2 V)^* D^2 U) - \scalar{\nabla \Delta U,\nabla V} ~. \end{align*} Recalling that $D^2 U \geq 0$ and $D^2 V \geq 0$, we conclude that: \begin{equation} \label{eq:Z-PDE} \frac{d}{dt} Z \leq \Delta Z - 2 \scalar{\nabla Z , \nabla V} - \scalar{\nabla Z,\nabla U} - \scalar{\nabla \Delta U,\nabla V} ~. \end{equation} To apply the maximum principle to (\ref{eq:Z-PDE}), we need to control the zeroth order (right-most) term. To this end, we claim that: \begin{equation} \label{eq:DV-bound} \norm{\nabla V_t}_{L_\infty} \leq \norm{\nabla V_0}_{L_\infty} \;\;\; \forall t \geq 0 ~. \end{equation} This follows e.g. by using the pointwise estimate of Bakry and \'Emery, refined by Bakry \cite[Proposition 1]{BakryRieszPartII}, which when $U$ is convex yields $|\nabla P_t^U(f)| \leq P_t^U(|\nabla f|)$. Together with the maximum principle, this indeed implies that: \[ |\nabla V_t(x)| = \frac{|\nabla P_t^U(\exp(-V_0))(x)|}{P_t^U(\exp(-V_0))(x)} \leq \frac{P_t^U(|\nabla V_0|\exp(-V_0))(x)}{P_t^U(\exp(-V_0))(x)} \leq \norm{\nabla V_0}_{L_\infty} ~. \] Now applying formally the maximum principle to (\ref{eq:Z-PDE}), using (\ref{eq:DV-bound}) and recalling the definition of $Z$, we obtain: \[ \norm{\Delta V_t}_{L_\infty} \leq \norm{\Delta V_0}_{L_\infty} + t\, n \norm{D^3 U}_{L_\infty} \norm{D V_0}_{L_\infty} ~. \] The assumption (\ref{eq:sec3-assumptions}) ensures (in particular) that all terms above are bounded, and hence $\Delta V_t$ is uniformly bounded on $[0,T]$ and it seems that we are done. However, there is a technical issue here: to appeal to the maximum principle on the unbounded domain $\Real^n$, we have to {\em a-priori} verify that $\Delta V_t (x)$ does not grow spatially faster than $\exp(C |x|^2)$ for some $C>0$, uniformly in $t \in [0,T]$ (see e.g. \cite[Chapter 2, Theorem 9]{FriedmanPDEBook}). The rest of the proof is dedicated to verifying this a-priori growth rate. \bigskip First, observe that $V$ grows spatially at most linearly, uniformly in $t \in [0,T]$. To see this without eluding to compactness, denote by $m$ the minimum of $V_0$, and hence (by the maximum principle) of $V(\cdot,t)$ for any $t \geq 0$. Fix $C>0$ and let $r>0$ be so that: \[ \exp(-C) \mu(B(r)) + \exp(-m) (1 - \mu(B(r))) < 1 ~. \] It follows since $\int \exp(-V(x,t)) d\mu(x) = 1$ for any $t \geq 0$, that for any such $t$ there exists $x_0(t) \in B(r)$ so that $V(x_0(t),t) \leq C$. Consequently, (\ref{eq:DV-bound}) implies that $V(x,t) \leq \norm{\nabla V_0}_{L_\infty} |x -x_0(t)| + V(x_0(t),t) \leq \norm{\nabla V_0}_{L_\infty} (|x|+r) + C$. Now write (\ref{eq:V-PDE}) as: \[ \frac{d}{dt} V - \Delta V = h ~,~ V|_{t=0} = V_0 ~, \] where $-h = \scalar{\nabla V,\nabla V} + \scalar{\nabla V,\nabla U}$. By the assumptions of Theorem \ref{thm:main1}, $|\nabla U|(x)$ grows at most linearly in $|x|$, and together with (\ref{eq:DV-bound}), it follows that $h$ too grows spatially at most linearly. Consequently, applying an interior regularity estimate (e.g. applying the estimate \cite[Chapter IV, (10.2)]{LadyParabolicBook} for the Sobolev space $W_p^{(2;1)}$ with $p$ arbitrarily large, followed by a variant of Morrey's embedding theorem as in the Corollary after \cite[Chapter IV,Theorem 9.1]{LadyParabolicBook}), it follows that: \begin{eqnarray*} & & \norm{V}_{C^{(1+\alpha;(1+\alpha)/2)}(B(R)\times[0,T])} \\ & \leq & C(n,T,\alpha) ( \norm{h}_{C^0(B(R')\times[0,T])} + \norm{V_0}_{C^2(B(R'))} + \norm{V}_{C^{0}(B(R') \times [0,T])}) ~, \end{eqnarray*} for any $\alpha \in (0,1)$ and $1 \leq R \leq R'-1$. Since $\norm{\nabla V_0}_{C^{1+\alpha}(\Real^n)}$ is assumed bounded in (\ref{eq:sec3-assumptions}), and as explained above, $V_0$, $V$ and $h$ grow spatially at most linearly, it follows that so does $\norm{V}_{C^{(1+\alpha;(1+\alpha)/2)}(B(R)\times[0,T])}$. Using this and arguing as above, we verify that $\norm{h}_{C^{(\alpha;\alpha/2)}(B(R)\times[0,T])}$ grows at most quadratically in $R$. Applying the interior Schauder estimate again (e.g. \cite[Chapter IV, Theorem 10.1]{LadyParabolicBook}), it follows that: \begin{eqnarray*} & & \norm{V}_{C^{(2+\alpha;1+\alpha/2)}(B(R)\times[0,T])} \\ & \leq & C(n,T,\alpha) ( \norm{h}_{C^{(\alpha;\alpha/2)}(B(R')\times[0,T])} + \norm{V_0}_{C^{2+\alpha}(B(R'))} + \norm{V}_{C^0(B(R')\times[0,T])} ) ~, \end{eqnarray*} for any $1 \leq R \leq R'-1$. Using (\ref{eq:sec3-assumptions}) again, we conclude that $D^2 V_t$ \emph{a-priori} spatially grows at most polynomially, uniformly in $t \in [0,T]$, thereby concluding the proof. \end{proof} We conclude from Lemma~\ref{lem:hessian of V bound} that the maps $S_t$ are well-defined (at least under the assumption (\ref{eq:sec3-assumptions})). Moreover, it follows that $S_t$ are diffeomorphisms (e.g. \cite[Theorem 1.61]{GHLBookEdition2}), since the inverse maps $T_{t,t} = T_t := S_t^{-1}$ may be obtained by running the flow backwards: \[ \frac{d}{d\tau} T_{t,\tau}(x) = -W_{t-\tau}(T_{t,\tau}(x)) ~,~ T_{t,0} = Id ~,~ \tau \in [0,t] ~. \] Clearly, the maps $S_t$ and $T_t$ inherit the symmetries of the vector field $W_t = - \nabla \log P_t^U(\exp(-V))$. As explained in the proof of Theorem \ref{thm:max}, $- \log P_t^U(\exp(-V))$ is invariant under the common symmetries of $U$ and $V$, i.e. our symmetry assumptions (\ref{eq:symmetry}), and so its gradient commutes with the group $O(E_1,\ldots,E_k)$ ; our maps therefore satisfy our symmetry assumptions as well. Theorem \ref{thm:main2} guarantees that $D W_t \geq 0$ and hence $(D W_t)^* + D W_t\geq 0$ for every $t \geq 0$. Consequently: \[ \frac{d}{dt} (DS_t)^*(x) DS_t(x) = (DS_t)^*(x) (D W_t)^*(S_t x) DS_t(x) + (DS_t)^*(x) D W_t(S_t x) DS_t(x) \geq 0 ~, \] and hence $(DS_t)^* DS_t \geq Id$ for every $t \geq 0$. In other words, $S_t$ is \emph{locally} an expansion. Since $S_t$ is also a diffeomorphism, it follows that it is in fact an expansion \emph{globally}. Indeed, $(DT_t)^* DT_t \leq Id$, which implies by integration and the triangle inequality that $|T_t(x) - T_t(y)| \leq |x-y|$. \medskip Next, we address the question of convergence of $\nu_t := P_t^U(\exp(-V)) \mu$ to $\mu$. Although we will only require convergence in $L_1$ for the sequel, we state the following for completeness: \begin{lem} \label{lem:convergence} As $t \rightarrow \infty$, we have: \begin{enumerate} \item $P_t^U(\exp(-V)) \rightarrow 1$ in $L_p(\mu)$, for any $p \in [1,\infty)$. \item $P_t^U(\exp(-V)) \rightarrow 1$ in $L_\infty(C)$, for any compact set $C \subset \Real^n$. \item $\norm{\frac{d\nu_t}{dx} - \frac{d\mu}{dx}}_{L_p} \rightarrow 0$ for any $p \in [1,\infty]$. \end{enumerate} \end{lem} For the proof, first recall that by (\ref{eq:L-prop}), $-L = -\Delta + \scalar{\nabla,\nabla U}$ is a symmetric positive semi-definite operator on the subspace $C^\infty(\Real^n) \cap L_2(\mu)$, and hence admits a Friedrichs extension to a self-adjoint positive semi-definite operator on a larger dense subspace $\D$ of $L_2(\mu)$, which we also denote by $-L$. Since $U$ is convex and $\mu = \exp(-U(x)) dx$ is a probability measure, it is known that $-L$ has a strictly positive spectral-gap $\lambda_1 > 0$ away from the trivial eigenvalue of $0$, corresponding to the constant functions: $\int - f L f d\mu \geq \lambda_1 \int f^2 d\mu$ for all $f \in \D_0 := \set{ f \in \D ; \int f d\mu = 0}$. For instance, by \cite{KLS} (see also \cite{EMilman-RoleOfConvexity}), one may estimate $\lambda_1 \geq c (\int |x| d\mu(x))^{-2} > 0$ for some universal numeric constant $c>0$. \begin{proof}[Proof of Lemma~\ref{lem:convergence}] Since $\lambda_1$ is strictly positive, the Spectral Theorem implies that $P_t^U(\exp(-V)) = \exp(-tL)(\exp(-V))$ tends in $L_2(\mu)$ to the projection of $\exp(-V)$ onto the constant functions, i.e. to the constant function $1 = \int \exp(-V) d\mu$. Since $P_t^U$ is bounded in $L_\infty$ (as in Subsection \ref{subsec:smooth}), we deduce the first claim for $p \in [2,\infty)$ by interpolation (and by Jensen's inequality this extends to $p \in [1,\infty)$). Next, we follow an argument similar to that used by Ledoux \cite{LedouxLogSobLectureNotes}. Denoting $f = \exp(-V)$, write: \[ |P_t^U(f)(x) - 1| = |P_t^U(f)(x) - \int P_t^U(f)(y) d\mu(y)| \leq \int |P_t^U(f)(x) - P_t^U(f)(y)| d\mu(y) ~. \] Certainly $|P_t^U(f)(x) - P_t^U(f)(y)| \leq |\grad P_t^U(f)(z)| |x-y|$ for some intermediate point $z \in [x,y]$. But using that $U$ is convex, the following smoothing estimate is known (\cite{LedouxSpectralGapAndGeometry}): \[ |\grad P_t^U(f)(z)| \leq \frac{1}{\sqrt{2t}} \norm{f}_{L_\infty} ~, \] and so: \[ |P_t^U(f)(x) - 1| \leq \frac{1}{\sqrt{2t}} \norm{f}_{L_\infty} \brac{|x| + \int |y| d\mu(y)} ~. \] The uniform convergence (as $t \to \infty$) on compact subsets follows. Moreover, since $|x| \exp(-U(x))$ is necessarily bounded, we obtain the third claim for $p = \infty$. The third claim for $p=1$ is equivalent to the first one with $p=1$, and so by interpolation, the third claim follows for all $p \in [1,\infty]$. \end{proof} Recall that a sequence of Borel measures $\set{\eta_k}$ is said to converge to a Borel measure $\eta$ \emph{weakly} (or in the weak$^*$-topology) if $\int \varphi d\eta_k \rightarrow \int \varphi d\eta$ for any bounded continuous test function $\varphi$; we will denote this by $\eta_k \rightharpoonup \eta$. We define the $L_1$ distance between two absolutely continuous Borel measures $\eta_1,\eta_2$ on $\Real^n$ to be: \[ d_{L_1}(\eta_1,\eta_2) := \int_{\Real^n} \abs{\frac{d\eta_1}{dx} - \frac{d\eta_2}{dx}} dx ~; \] this coincides with the usual total-variation distance up to a factor of $2$. Clearly, convergence in $L_1$ implies weak convergence. \begin{lem} \label{lem:approx} Let $\set{\mu_k}$ and $\set{\nu_k}$ denote two sequences of absolutely continuous Borel measures on $\Real^n$, such that each $\nu_k$ is the push-forward of $\mu_k$ by a contracting map $T_k : \Real^n \rightarrow \Real^n$. Assume that $d_{L_1}(\mu_k,\mu) \rightarrow 0$ and $\nu_k \rightharpoonup \nu$. Then there exists a contraction $T : \Real^n \rightarrow \Real^n$ pushing forward $\mu$ onto $\nu$. Moreover, any common symmetries possessed by $T_k$ are preserved by $T$. \end{lem} \begin{proof} First, note that $T_k(0)$ must be uniformly bounded. Indeed, let $B(R_1)$ denote a ball around the origin so that $\mu(B(R_1)) \geq 3/4$. The $L_1$ convergence immediately implies that $\mu_k(B(R_1)) \rightarrow \mu(B(R_1))$, and so $\mu_k(B(R_1)) \geq 2/3$ for large enough $k$. Similarly, if $B(R_2)$ denotes a ball so that $\nu(B(R_2)) \geq 3/4$, it follows easily from the weak convergence that $\nu_k(B(0,R_2)) \rightarrow \nu(B(R_2))$ (here we need to use the fact that the ball has finite perimeter and that our measures are absolutely continuous with respect to Lebesgue measure), and hence $\nu_k(B(R_2)) \geq 2/3$ for large enough $k$. Consequently, for large enough $k$, $\mu_k(T_k^{-1}(B(R_2))) = \nu_k(B(R_2)) \geq 2/3$, and therefore $T_k^{-1}(B(R_2)) \cap B(R_1)$ is non-empty. Since $T_k$ is a contraction, it follows that $T_k(0) \in B(R_1+R_2)$. Next, by passing to a subsequence if necessary, we may assume that $T_k(0)$ converges. Since $T_k$ are all contractions, and hence uniformly (Lipschitz) continuous, it follows by compactness and a standard diagonalization argument that, after passing to an appropriate subsequence, $T_k$ uniformly converges on compact subsets of $\Real^n$ to some map $T$, which is consequently a contraction, which preserves the common symmetries of $T_k$. It remains to show that $T$ pushes forward $\mu$ onto $\nu$. This is equivalent to showing that $\int \varphi(Tx) d\mu(x) = \int \varphi(y) d\nu(y)$ for any bounded continuous test function $\varphi : \Real^n \rightarrow \Real$. Since by definition, for any $k$: \[ \int \varphi(T_k x) d\mu_k(x) = \int \varphi(y) d\nu_k(y) ~, \] and the right hand side converges to $\int \varphi(y) d\nu(y)$, it remains to show that the left hand side converges to $\int \varphi(Tx) d\mu(x)$. Indeed: \begin{multline*} \abs{ \int \varphi(T_k x) d\mu_k(x) - \int \varphi(Tx) d\mu(x) } \\ \leq \abs{ \int \varphi(T_k x) d\mu_k(x) - \int \varphi(T_k x) d\mu(x) } + \abs{ \int \varphi(T_k x) d\mu(x) - \int \varphi(T x) d\mu(x) } ~. \end{multline*} The first term on the right hand side converges to $0$ since $\varphi$ is bounded and $d_{L_1}(\mu_k,\mu) \rightarrow 0$. The second term converges to $0$ by Lebesgue's dominant convergence theorem, since (the bounded) $\varphi(T_k x)$ pointwise converges to $\varphi(Tx)$ (in fact uniformly on compact subsets). This concludes the proof. \end{proof} Lemma \ref{lem:convergence} (case (3) with $p=1$) ensures that $\nu_t$ converges to $\mu$ in $L_1$. Since $\nu$ is the push-forward of $\nu_t$ via $T_t$ which is a contraction, it follows by Lemma \ref{lem:approx} that there exists a contraction $T_\infty$ pushing forward $\mu$ onto $\nu$ and satisfying our symmetry assumptions. This concludes the proof of Theorem \ref{thm:main1} in the case that $U$ and $V$ are assumed smooth and under the additional assumptions of (\ref{eq:sec3-assumptions}). To conclude the theorem in the full generality, apply Lemma \ref{lem:approx} again to see that there exists a contraction pushing forward $\mu$ onto $\nu$, whenever these measures may be approximated by smooth measures satisfying the assumptions of the theorem and (\ref{eq:sec3-assumptions}). Such approximation is always possible by a standard argument: applying the Legendre transform to $V$, redefining the resulting function to be $+\infty$ beyond some large level, and applying the transform again, we obtain a convex Lipschitz function with the same symmetries, and it remains to convolve it with a smooth rotation-invariant mollifier, yielding the first part of (\ref{eq:sec3-assumptions}) ; a similar argument applies to the function $U$, whose special form \eqref{eq:U} reduces the approximation to an easy one-dimensional problem. Lemma \ref{lem:approx} thus implies the general case of Theorem \ref{thm:main1}. \section{Applications} \label{sec:apps} The first application we would like to describe pertains to a generalization of the Gaussian Correlation Conjecture. This conjecture asks whether for any two convex subsets $A,B \subset \Real^n$, which are in addition centrally-symmetric ($C$ is called centrally-symmetric if $C = -C$), the following inequality is valid for the standard Gaussian measure $\gamma_n$ on $\Real^n$: \begin{equation} \label{eq:GCC} \gamma_n(A \cap B) \geq \gamma_n(A) \gamma_n(B) ~? \end{equation} We refer to \cite{SSZ-GaussianCorrelationConjecture,HargeGCCForEllipsoid,CorderoMassTransportAndGaussianInqs} and the references therein for the history of this conjecture, which remains open for $n \geq 3$. One of the most general results is due to Harg\'e \cite{HargeGCCForEllipsoid}, who confirmed the validity of (\ref{eq:GCC}) when one of the sets is a (centrally-symmetric) ellipsoid. This was subsequently given a different proof by Cordero-Erausquin \cite{CorderoMassTransportAndGaussianInqs}, as a direct corollary of Caffarelli's Contraction Theorem (in this context, it is worthwhile pointing out that our construction of the expanding map $T^{-1}$ closely resembles Harg\'e's argument). Replacing Caffarelli's theorem with Theorem \ref{thm:main1} in Cordero-Erausquin's argument, we obtain the following generalization: \begin{cor} \label{cor:GCC} Let $\mu = \exp(-U(x)) dx$ denote a probability measure on $\Real^n$ as in Theorem \ref{thm:main1}, which is in addition centrally symmetric (i.e. the quadratic part of $U$ on $E_0$ is assumed even). Let $B$ denote a centrally-symmetric \emph{convex} subset of $\Real^n$ satisfying the following symmetry assumptions: \[ \exists C_B \subset \Real^{dim E_0 + k} \;\;\;\; \mathbf{1}_B(x) = \mathbf{1}_{C_B}(Proj_{E_0} x,|Proj_{E_1} x|,\ldots,|Proj_{E_k} x|) ~. \] Let $A$ denote a centrally-symmetric subset of $\Real^n$ so that, writing for $x \in \Real^n$, $x = (x_0,x_1,\ldots,x_k)$ with $x_i \in E_i$, we have: \begin{multline} \label{eq:A-prop} \text{ if } (x_0,x_1,\ldots,x_k) \in A \text{ then } \\ \forall y_0 \in E_0, \;\; \norm{y_0}_{\E} \leq \norm{x_0}_{\E}, \; \forall t_i \in [-1,1], \text{ we have } (y_0, t_1 x_1, \ldots, t_k x_k) \in A ~, \end{multline} where $\norm{\cdot}_{\E}$ is the norm associated with some centrally-symmetric ellipsoid $\E \subset E_0$. Then: \[ \mu(A \cap B) \geq \mu(A) \mu(B) ~. \] \end{cor} Clearly, this generalizes the result of Harg\'e and Cordero--Erausquin, by choosing $\mu = \gamma_n$ and $E_0 = \Real^n$. \begin{proof} First, by applying an appropriate linear transformation $P$ in $E_0$ which leaves the orthogonal complement invariant, we may assume that $\E$ is a Euclidean ball in $E_0$, since $P(B)$ and $P_*(\mu)$ continue to satisfy the assumptions of the theorem (indeed, $P$ only affects the quadratic part of $U$, which remains quadratic and even). Defining the probability measure $\mu_B$ as the restriction of $\mu$ onto $B$, i.e. $\mu_B(C) = \mu(C \cap B) / \mu(B)$, our task is to show that $\mu_B(A) \geq \mu(A)$. It is standard to approximate $\mathbf{1}_B / \mu(B)$ in $L_1(\Real^n)$ by functions of the form $\exp(-V_k)$, where $V_k$ is convex and satisfies the same symmetries as $B$, implying that $\exp(-V_k) \mu$ tends to $\mu_B$ in total-variation. Applying Theorem \ref{thm:main1} and Lemma \ref{lem:approx}, we deduce that there exists a contraction $T$ pushing forward $\mu$ onto $\mu_B$ and satisfying our symmetry assumptions. Since $T$ commutes with $O(E_1,\ldots,E_k)$, it follows easily that $Proj_{E_i} T(x)$ is radial for $i=1,\ldots,k$: \begin{equation} \label{eq:radial-sym} Proj_{E_i} T(x) = T_i(x_0,|x_1|,\ldots,|x_k|) \frac{x_i}{|x_i|} \text{ if $x_i \neq 0$ and 0 otherwise} ~. \end{equation} Moreover, since $B$ and $\mu$ were assumed centrally-symmetric, it is easy to check that $T$ will also preserve this additional symmetry. Denoting by $R_i$ the reflection in the subspace $E_i$, i.e. $R_i(x) = x - 2Proj_{E_i} x$ for $i=0,1,\ldots,k$, we conclude that $T$ commutes with all the $R_i$'s. It remains to note that $T(A) \subset A$. Indeed, using the commutation with $R_i$ and the contraction property of $T$, we have: \[ 2 |Proj_{E_i} T(x)| = |R_i (T(x)) - T(x)| = | T(R_i(x)) - T(x) | \leq |R_i(x) - x| = 2 |Proj_{E_i} x| ~, \] and so $|Proj_{E_i} T(x)| \leq |Proj_{E_i} x|$ for $i=0,1,\ldots,k$. Together with (\ref{eq:radial-sym}) and the symmetries (\ref{eq:A-prop}) of $A$, it follows that $T(A) \subset A$. Consequently $A \subset T^{-1}(A)$, and therefore: \[ \frac{\mu(A \cap B)}{\mu(B)} = \mu_B(A) = \mu(T^{-1}(A)) \geq \mu(A) ~. \] The proof is complete. \end{proof} \begin{rem} It is possible to replace the requirement $t_i \in [-1,1]$ in (\ref{eq:A-prop}) by $t_i \in [0,1]$. This is achieved by using the Brenier map $T_{opt}$ of Theorem \ref{thm:main3} instead of $T$ in the proof above, thereby ensuring that the $\set{T_i}_{i=1}^k$ in (\ref{eq:radial-sym}) are always non-negative, as explained in Section \ref{sec:revisit}. \end{rem} The following two additional corollaries may be easily obtained from the previous one by integration by parts: \begin{cor} Let $\mu$ denote a probability measure on $\Real^n$ as in Theorem \ref{thm:main1}, which is in addition centrally symmetric. Let $f,g : \Real^n \rightarrow \Real_+$ denote two measurable bounded functions, so that for each $a,b > 0$, the level sets $f^{-1}([a,\infty))$ and $g^{-1}([b,\infty))$ satisfy the assumptions on the sets $A$ and $B$ in Corollary \ref{cor:GCC}, respectively. Then: \[ \int f g d\mu \geq \int f d\mu \int g d\mu ~. \] \end{cor} \begin{cor} Let $\mu,\nu$ denote two probability measures as in Theorem \ref{thm:main1}, and assume in addition that both are centrally symmetric. Let $\Gamma : \Real^n \rightarrow \Real_+$ denote a measurable function such that all of its level sets $\Gamma^{-1}([0,a])$ (individually) satisfy the assumption on the set $A$ in Corollary \ref{cor:GCC}. Then: \[ \int \Gamma(x) d\nu(x) \leq \int \Gamma(x) d\mu(x) ~. \] \end{cor} These corollaries generalize the correlation inequalities obtained in \cite{BrascampLiebPLandLambda1,HargeGCCForEllipsoid,CaffarelliContraction} for the case $dim E_0 = n$. We remark that when $dim E_0 = 0$, the corollaries may be obtained directly without appealing to Theorem \ref{thm:main1}, so the more interesting case is when $0 < dim E_0 < n$. \medskip Finally, we also mention that contracting maps constitute a very useful tool to transfer isoperimetric inequalities from one measure-metric space to another. Note that the measure $\mu$ of Theorem \ref{thm:main1} is a product measure, with each factor being either a Gaussian or a log-concave radially symmetric measure. The isoperimetric inequality satisfied by the former factor is well known \cite{SudakovTsirelson,Borell-GaussianIsoperimetry}, and has recently been identified (up to numeric constants) for the latter factor \cite{HuetSphericallySymmetric}. The tools to transfer these inequalities to the product measure have also recently been obtained \cite{BartheTensorizationGAFA,BCRHard,BCRSoft,EMilmanRoleOfConvexityInFunctionalInqs}, and so consequently, the isoperimetric inequality satisfied by $\mu$ is well understood. Using the contracting map $T$ of Theorem \ref{thm:main1}, it follows that the same isoperimetric inequality is satisfied by the measure $\nu$. We refer to \cite{LatalaJacobInfConvolution} for further examples of using contracting maps to transfer isoperimetric inequalities, and for further information. \section{Caffarelli's proof revisited} \label{sec:revisit} Let us now sketch the proof of Theorem \ref{thm:main3}, which is based on the proof of \cite[Theorem 11]{CaffarelliContraction}, but requires an additional ingredient from \cite{CaffarelliContraction} in the form of Theorem \ref{thm:ingred} below. Throughout this section we use $T$ to denote the Brenier optimal-map. \subsection{The Radial Case} We begin with the elementary case when $\mu = \exp(-\rho(|x|)) dx$ and $\nu = \exp(-(\rho+v)(|x|)) dx$ are radial. This case does not require the use of Theorem \ref{thm:ingred}, and as we will see, clearly motivates the condition $\rho''' \leq 0$ in Theorem \ref{thm:main1}. First, it is immediate to reduce to the one dimensional case, when $\mu$ and $\nu$ are supported on $\Real_+$. Indeed, by the radial symmetry and the uniqueness of the Brenier map $T = \nabla \varphi$ with $\varphi: \Real^n \rightarrow \Real$ a convex function, it follows that $T$ must also be radially symmetric, i.e. commute with the orthogonal group. Consequently, we may write $\varphi(x) = \phi(|x|)$ with $\phi: \Real_+ \rightarrow \Real$ convex, and $T(r \theta) = T_1(r) \theta$ for $\theta \in S^{n-1}$ and $r \in \Real_+$. $T_1 = \phi' : \Real_+ \rightarrow \Real_+$ is precisely the Brenier map pushing forward $\exp(-\rho(r)) r^{n-1} dr$ onto $\exp(-(\rho(r)+v(r))) r^{n-1} dr$. Denoting $\rho_1(r) = \rho(r) - (n-1) \log r$, we see that $\rho_1$ remains convex and $\rho_1''' \leq 0$, and so it is enough to show that when in addition $v : \Real_+ \rightarrow \Real$ is convex and non-decreasing, the Brenier map $T_1$ pushing forward $\mu_1 = \exp(-\rho_1(r)) dr$ onto $\nu_1 =\exp(-(\rho_1(r)+v(r))) dr$ is a contraction. Indeed, in the one dimensional case, the derivative of a convex function is simply a monotone non-decreasing one, and so the Brenier map is the unique non-decreasing map pushing forward $\mu_1$ onto $\nu_1$, given by: \begin{equation} \label{eq:1D-pf} \int_0^{T_1(x)} \exp(-(\rho(r)+v(r))) dr = \int_0^x \exp(-\rho(r)) dr ~. \end{equation} Since $\rho,v$ are assumed smooth enough, so is $T_1$. Taking derivatives, we obtain: \begin{equation} \label{eq:1D-MA} \log T_1'(x) = -\rho(x) + \rho(T_1(x)) + v(T_1(x)) ~. \end{equation} Assume that the maximum of $T_1'$ is attained at $x_0 \in \Real_+$. To ensure this, one would actually need to restrict $\nu_1$ onto a compact subset, in which case $\lim_{x \rightarrow \infty} T_1'(x) = 0$ and so the (positive) maximum is attained, and conclude with an approximation argument (as in \cite{CaffarelliContraction}) ; we omit the details here. Our task is to show that $T_1'(x_0) \leq 1$. If $x_0 = 0$, since $T_1(0) = 0$ and $\exp(-v(0)) \geq 1$ (otherwise $\mu$ and $\nu$ could not both have total mass 1), it follows that $T_1'(0) \leq 1$, as required. Otherwise, denoting $F = \log T_1'$, since $F$ and $T_1'$ have a local maximum at $x_0$, it follows that $T''_1(x_0) = 0$ and that: \begin{eqnarray*} \!\!\!\!\!\!\!\! 0 & \geq & F''(x_0) \\ & = & -\rho_1''(x_0) + (T_1'(x_0))^2 (\rho_1''(T_1(x_0)) + v''(T_1(x_0))) + T_1''(x_0)(\rho_1'(T_1(x_0)) + v'(T_1(x_0))) \\ & = & -\rho_1''(x_0) + (T_1'(x_0))^2 (\rho_1''(T_1(x_0)) + v''(T_1(x_0))) ~. \end{eqnarray*} Since $v'' \geq 0$ and $\rho_1'' \geq 0$, we obtain that: \begin{equation} \label{eq:compare} (T_1'(x_0))^2 \leq \frac{\rho_1''(x_0)}{\rho_1''(T_1(x_0))} ~. \end{equation} In Caffarelli's argument, $\rho_1$ is a quadratic polynomial, and therefore the right-hand side above is identically $1$. However, since $T_1(x) \leq x$ for all $x \in \Real_+$, as easily verified from (\ref{eq:1D-pf}) and the fact that $v$ is non-decreasing, we obtain by the mean-value theorem that the right-hand side is not greater than 1 as soon as $\rho_1''' \leq 0$. This concludes the proof and explains the latter condition. \medskip We remark that in this simple case, the Brenier map and the map we construct in our proof of Theorem \ref{thm:main1} do in fact coincide, since the latter one is also radially symmetric, and is constructed as a limit of diffeomorphisms, and hence must be monotone on each ray from the origin. \subsection{The General Case} Let $\mu = \exp(-U(x)) dx$ and $\nu = \exp(-(U(x) + V(x))) dx$ be two probability measures in $\Real^n$, satisfying the assumptions in Theorem \ref{thm:main1}. We will actually assume that $\nu$ is supported on a compact convex set $C$, to be specified later on, and that $U \in C^{3,\alpha}(\Real^n)$, $U$ is strictly convex, and $V \in C^{3,\alpha}(C)$ ; the general case follows by a standard approximation argument, under which one may show that the corresponding Brenier maps converge to the gradient of a convex function, i.e. the Brenier map for the limiting measures, and the contraction property is trivially preserved in the limit. Let $T = \nabla \varphi$ denote the Brenier map pushing forward $\mu$ onto $\nu$, where $\varphi : \Real^n \rightarrow \Real$ is a convex potential. It follows from our assumptions and Caffarelli's regularity theory \cite{CaffarelliStrictlyConvexIsHolder,CaffarelliHigherHolderRegularity,CaffarelliRegularity} that $\varphi \in C^{5,\alpha}_{loc}(\Real^n)$. It also follows from the proof of \cite[Lemma 4]{CaffarelliContraction} and the subsequent remark that $\norm{D T}(x) = \max_{\xi \in S^{n-1}} D^2_{\xi,\xi} \varphi(x)$ attains a maximum in $\Real^n$, since $D^2_{\xi,\xi} \varphi(x)$ tends to $0$ as $|x| \rightarrow \infty$ uniformly in $\xi \in S^{n-1}$, when $C$ is convex. We will denote by $x_0$ a point where this maximum is attained. Our task is to show that $\norm{T}_{Lip} := D^2_{e,e} \varphi(x_0) \leq 1$, where $e \in S^{n-1}$ is the eigenvector of $D^2 \varphi(x_0)$ corresponding to its maximal eigenvalue, and hence: \begin{equation} \label{eq:e-max} D_e D \varphi(x_0) = D^2_{e,e} \varphi(x_0) e ~. \end{equation} As usual, attaining the maximum at $x_0$ implies that: \begin{equation} \label{eq:MA-max} \nabla D^2_{e,e} \varphi(x_0) = D^2_{e,e} T(x_0) = 0 ~ , ~ D^2 D^2_{e,e} \varphi(x_0) = D^2_{e,e} DT(x_0) \leq 0 ~. \end{equation} As in (\ref{eq:1D-MA}), the change-of-variables formula resulting from the definition of push-forward is: \begin{equation} \label{eq:MA} \log \det DT(x) = -U(x) + U(T(x)) + V(T(x)) ~. \end{equation} Differentiating (\ref{eq:MA}) twice in the direction of $e$, we obtain: \begin{eqnarray} \label{eq:bigf} \!\!\!\!\!\! & & - tr( (DT)^{-1}(x) D_e DT(x) (DT)^{-1}(x) D_e DT(x) ) + tr( (DT)^{-1}(x) D^2_{e,e} DT(x) )\\ \nonumber \!\!\!\!\!\! & = & -D^2_{e,e} U(x) + \scalar{ D^2 (U+V) (T(x)) D_e T(x), D_e T(x)} + \scalar{D (U+V)(T(x)), D^2_{e,e} T(x) } ~. \end{eqnarray} Using that $DT = D^2 \varphi > 0$, observe that $D_e DT (DT)^{-1} D_e DT \geq 0$. Recalling by (\ref{eq:MA-max}) that $D^2_{e,e} DT(x_0) \leq 0$, and using the fact that $tr(AB) \geq 0$ if $A,B \geq 0$, it follows that the left-hand side of (\ref{eq:bigf}) is non-positive when evaluated at $x_0$. Noting by (\ref{eq:MA-max}) that the last summand on the right-hand side of (\ref{eq:bigf}) vanishes at this point, and using $D^2 V \geq 0$ and (\ref{eq:e-max}), we conclude that: \[ D^2_{e,e} U(x_0) \geq \scalar{ D^2 U(T(x_0)) D_e D \varphi (x_0), D_e D \varphi (x_0) } = D^2_{e,e} U(T(x_0)) |D^2_{e,e} \varphi(x_0)|^2 ~. \] Since $D^2 U > 0$, we obtain the analogue of (\ref{eq:compare}): \[ \norm{T}_{Lip}^2 = |D^2_{e,e} \varphi(x_0)|^2 \leq \frac{D^2_{e,e} U(x_0)}{D^2_{e,e} U(T(x_0))} ~. \] When $U$ is quadratic, this is already enough to guarantee that $T$ is contracting. To make sure that the right-hand side is not greater than 1 under more general circumstances, we would need by the mean-value theorem to ensure that: \begin{equation} \label{eq:ensure} \left . (D^3 U) \right |_{y}(e,e,x_0 - T(x_0)) \leq 0 \;\;\; \forall y \in [x_0,T(x_0)] ~. \end{equation} By the uniqueness of the Brenier map and the symmetries of $\mu$ and $\nu$, we know that $T$ must satisfy our symmetry assumptions. Consequently, as in the proof of Corollary \ref{cor:GCC}, $T$ must act radially on each $E_i$, $i=1,\ldots,k$: \[ Proj_{E_i} T(x) =\begin{cases} T_i(Proj_{E_0} x,|Proj_{E_1} x|,\ldots,|Proj_{E_k} x|) \frac{Proj_{E_i} x}{|Proj_{E_i} x|} & \text{ if $Proj_{E_i} x \neq 0$}, \\ 0 & \text{otherwise}. \end{cases} \] As the gradient of a convex function, we must have $\scalar{T(x) - T(y) , x - y} \geq 0$ for all $x,y \in \Real^n$, and using $y = x - 2 Proj_{E_i} x$ (reflecting $x$ in $E_i$ about the origin) implies that necessarily $T_i \geq 0$. Consequently: \[ \forall i=1,\ldots,k \;\;\; \exists a_i(x_0) \geq 0 \;\;\; Proj_{E_i} T(x_0) = a_i(x_0) Proj_{E_i} x_0 ~. \] We conclude from Lemma \ref{lem:geom1} that (\ref{eq:ensure}) would follow if we could show that: \begin{equation} \label{eq:a} \forall x \in \Real^n \;\;\; \forall i = 1,\ldots,k \;\;\; a_i(x) \leq 1 ~. \end{equation} Geometrically, this means we that we have reduced the task of showing that $T$ is a contraction, to showing that $T$ is a contraction \emph{with respect to the origin} on each $E_i$. Note that in the radial case, this followed trivially from the monotonicity of $v$. \medskip To show (\ref{eq:a}), we require the following additional ingredient \cite[Theorem 6]{CaffarelliContraction}. \begin{thm}[Caffarelli] \label{thm:ingred} Let $U_1 \in C^{1,\alpha}(\Omega_1)$ and $U_2 \in C^{1,\alpha}(\Omega_2)$, where $\Omega_2 = \times_{i=1}^n [a_i,b_i] \subset \Real^n$ and $\Omega_1 \supset \Omega_2$, so that $\int_{\Omega_i} \exp(-U_i(x)) dx = 1$. Let $\tilde{T}$ denote the Brenier optimal-transport map pushing forward $\exp(-U_1(x))dx$ onto $\exp(-U_2(x))dx$, and let $S$ denote a fixed subset of the coordinates $\set{1,\ldots,n}$. Assume that for any $x \in \Omega_1$, $y \in \Omega_2$ and $j \in S$: \begin{equation} \label{eq:ingred-cond} \text{$\forall i \in S \;\;\; y_i \leq x_i$ and $x_j = y_j$} \; \Rightarrow \; \frac{d}{dx_j} U_1 (x) \leq \frac{d}{dy_j} U_2(y) ~. \end{equation} Then $\tilde{T}(x)_i \leq x_i$ for all $i\in S$, for any $x \in \Omega_1$. \end{thm} In our formulation, we have exchanged between source and target measures (using that the Brenier map in this case is precisely the inverse of the original one), removed the assumption that $\Omega_1 = \Omega_2$, and consider only a subset of the coordinates for which the assumption and conclusion hold (as can be easily verified by inspecting the proof). \medskip Fix a coordinate structure determined by our decomposition of $\Real^n$ into $E_i$, let $Q$ denote the set of coordinates corresponding to $E_0$, and let $S$ denote the set of all other coordinates, corresponding to the subspaces $E_1,\ldots,E_k$. Set $C = [-R,R]^n$, $\Omega_1 = \Real^Q \times \Real_+^S$, $\Omega_2 = [-R,R]^Q \times [0,R]^S$, $U_1 = U + c_1$ and $U_2 = U + V + c_2$, where $c_i$ are constants designed to make $\exp(-U_i(x))dx$ probability measures on $\Omega_i$. The symmetries of $T$ described above imply that it is enough to verify (\ref{eq:a}) for $x \in \Omega_1$ and that $T|_{\Omega_1} = \tilde{T}$, where $\tilde{T}$ is given by Theorem \ref{thm:ingred}. Consequently, the desired (\ref{eq:a}) will follow from the conclusion of Theorem \ref{thm:ingred} if we verify (\ref{eq:ingred-cond}). Fix $j \in S$, corresponding to a subspace $E_l$. Lemma \ref{lem:geom2} implies that $\frac{d}{dy_j} V(y) \geq 0$ for any $y \in \Omega_2$, and so it is enough to verify that for $x \in \Omega_1$ and $y \in \Omega_2$: \begin{equation} \label{eq:final-red} \text{$\forall i \in S \;\;\; y_i \leq x_i$ and $x_j = y_j$} \; \Rightarrow \; \frac{d}{dx_j} U(x) \leq \frac{d}{dy_j} U(y) ~. \end{equation} But $\frac{d}{dx_j} U(x) = \frac{\rho_l'(|Proj_{E_l} x|)}{|Proj_{E_l} x|} x_j$, and when $x_j$ is fixed, the coefficient in front of it is non-increasing in $|Proj_{E_l} x|$ since $\rho_l'$ was assumed concave and $\rho_l'(0) = 0$. Since $x \in \Omega_1$ and $y \in \Omega_2$, the assumption $y_i \leq x_i$ for all $i \in S$ implies that $|Proj_{E_l} y| \leq |Proj_{E_l} x|$, confirming the desired (\ref{eq:final-red}). This finally concludes the proof. \section{Comparing the two maps} \label{sec:comparing} In this section, we compare the map $T$ (as constructed in Subsection~\ref{SS:construction of T}) with the Brenier map $T_{opt}$. First, it is natural to ask whether the two maps $T$ and $T_{opt}$ coincide, at least under the assumptions of Theorem~\ref{thm:main1}. To analyze this question, recall that $S_t$ was constructed as follows: \begin{equation} \label{eq:W-eq-again} \frac{d}{dt} S_t(x) = W_t(S_t(x)) ~,~ S_0 = Id ~,~ \text{with} ~~ W_t := \nabla Z_t ~,~ Z_t := -\log P_t^U(\exp(-V)) ~. \end{equation} Denoting $B_t(x) := D^2 Z_t(S_t(x))$ and taking spatial derivatives, we obtain: \begin{equation} \label{eq:DW-eq-again} \frac{d}{dt} DS_t(x) = B_t(x) DS_t(x) ~,~ DS_0(x) \equiv Id ~. \end{equation} As is well known, a necessary and sufficient condition for being the gradient of a function on a simply connected domain, is having a symmetric derivative tensor. It follows that: \begin{equation} \label{eq:commutation} \text{if all of $\set{B_t}_{t \geq 0}$ commute with each other} ~, \end{equation} ensuring that $DS_t$ remains symmetric along the flow, then we can conclude that $S_t$ is the gradient of some function (for each $t$). Moreover, we could then write: \[ DS_t(x) = \exp\brac{\int_0^t B_s(x) ds } ~, \] from which it would follow that $DS_t$ is pointwise positive semi-definite, and hence $S_t$ must be the gradient of a \emph{convex} function. The inverse map $T_t = S_t^{-1}$ would then be the gradient of a convex function as well, and this property may be shown to be preserved in the limit as $t \rightarrow \infty$, obtaining the Brenier map transporting $\mu = (S_\infty)_*(\nu)$ onto $\nu$. Condition (\ref{eq:commutation}) implies that in all one-dimensional situations ($n=1$ or radially symmetric data), both maps $T$ and $T_{opt}$ \emph{do} coincide. However, it is easy to check that generically, the sufficient condition (\ref{eq:commutation}) will be severely \emph{violated}, for instance by constructing examples (see below) so that for some $x$: \begin{equation} \label{eq:no-commutation2} 0 \neq [\frac{d}{dt} B_t(x), B_t(x) ] = [ D^2 \frac{d}{dt} Z_t + D^3 Z_t D Z_t, D^2 Z_t ](S_t(x)) \;\;\; \forall t \geq 0 ~, \end{equation} where $[A,B] = AB - BA$ denotes the Lie bracket. Moreover, it is not hard to show that (\ref{eq:no-commutation2}) implies that $DS_t(x)$ is non-symmetric on some non-empty interval $t \in (0,t_0)$, and that for any non-empty interval $(t_1,t_2) \subset (0,\infty)$, $\set{DS_t(x)}_{t \in (t_1,t_2)}$ cannot all commute. In other words, the path of diffeomorphisms $[0,\infty) \ni t \mapsto S_t$ will generically not coincide with the path of optimal interpolating maps $[0,1) \ni s \mapsto (1-s) Id + s S_{opt}$, where $S_{opt} = T_{opt}^{-1}$ is the Brenier map pushing forward $\nu$ onto $\mu$, and in fact the set of times $t$ where these two paths intersect will be discrete. All of this suggests that generically, the lack of symmetry (or path separation) should persist in the limit as $t \rightarrow \infty$, and hence that the limiting map $T$ should be different from $T_{opt}$. However, although we believe that (\ref{eq:commutation}) is actually also a necessary condition (at least generically) for obtaining the Brenier map, we are unable to rule out the possibility that the symmetry may be recovered in the limit. In particular, we are unable to show that the two maps are different even for the following simple example, where (almost) everything may be explicitly computed: \begin{example} Let $U,V$ be given by: \begin{align*} & U(x) = \frac{1}{2}\scalar{A x,x} ~,~ V(x) = \frac{1}{2}\scalar{B x,x} ~, \\ & \text{$A,B$ are positive-definite \emph{non-commuting} matrices} ~, \end{align*} and set: \[ \mu = c_1 \exp(-U(x))dx ~,~ \nu = c_2 \exp(-(U(x) + V(x))) dx ~, \] with $\set{c_i}$ chosen so that the resulting measures have total mass $1$. It is easy to see (e.g. \cite[Example 1.7]{McCannConvexityPrincipleForGases}) that $T_{opt}$ is a linear map given by the positive-definite matrix $C_{opt} = A^{1/2} (A^{1/2} (A+B) A^{1/2})^{-1/2} A^{1/2}$. The Mehler formula \cite{HargeGCCForEllipsoid} for an affine Ornstein-Uhlenbeck diffusion implies that the tensor $D^2 Z_t = -D^2 \log P_t^U(\exp(-V))$ is an explicitly computable fixed matrix $M_t$ for every time $t \geq 0$, and so by (\ref{eq:DW-eq-again}), the flow maps $\set{S_t}$ are also linear, given by a family of matrices $\set{L_t}$. Moreover, $L_t$ satisfy an explicit matrix-valued ODE, and one may also show that $L_t^* (A+M_t) L_t = A+B$. The resulting map $T$ is then the linear map given by the matrix $L_\infty^{-1}$, where $L_\infty = \lim_{t \rightarrow \infty} L_t$. Showing that $T \neq T_{opt}$ when $A,B$ do not commute then amounts to proving that $L_\infty$ is not symmetric in this case; we were unable to verify this. When $A,B$ do commute, then so do all the matrices $\set{M_t}$, so (\ref{eq:commutation}) is satisfied and $T = T_{opt}$. \end{example} An additional aspect of comparing between $T$ and $T_{opt}$ pertains to the condition that $U$ be convex in Theorems \ref{thm:main1} and \ref{thm:main3}. This condition was absolutely crucial in Caffarelli's argument and the proof of Theorem \ref{thm:main3}. However, an inspection of the proof of Theorem \ref{thm:main1} reveals that this condition was only used in the proof of (\ref{eq:Z-PDE}), (\ref{eq:DV-bound}) and Lemma \ref{lem:convergence}, and it is actually possible to relax our condition to $D^2 U \geq -c Id$ by a careful adaptation of the arguments (and in particular, avoid using the Spectral Theorem, since $-L$ will no longer have a spectral gap). Unfortunately, the convexity of $U$ actually follows from the other assumptions of Theorem \ref{thm:main1}, namely that $\mu = \exp(-U(x)) dx$ has finite total mass and that $\rho'''_i \leq 0$ on $\Real_+$, so ultimately there is no real gain here in using $T$ over $T_{opt}$. But this difference in the significance of the convexity of $U$ to the proof, perhaps reinforces the intuition that these two maps should be (generically) different. \medskip Before concluding, we mention a couple of advantages of working with the map $T$ over the Brenier map $T_{opt}$. In the proof of the contraction property of $T_{opt}$ (Theorem \ref{thm:main3}), Caffarelli's regularity theory for the fully-nonlinear Monge--Amp\`ere equation was an essential ingredient. In contrast, in our study of the map $T$ (Theorem \ref{thm:main1}), we only employed the classical regularity results for linear parabolic PDEs. This lends our heat-diffusion construction to further generalizations, in situations where the regularity for the Monge--Amp\`ere equation and the Brenier--McCann optimal-transport map has yet to be established, or alternatively is known to be false, for instance in the Riemannian-manifold setting (see \cite{VillaniOldAndNew}). In addition, other choices for the driving potential $Z_t$ in our flow scheme (\ref{eq:W-eq-again}) are also possible, in accordance to the property one wished to establish. \setlinespacing{0.93} \setlength{\bibspacing}{2pt} \bibliographystyle{plain} \def\cprime{$'$}
\section{\label{sec:introduction}Introduction} When suspensions of electrically-polarizable nanolayered clay particles suspended in silicon oil are subjected to an external electric field of the order of $\sim 1$ kV/mm, the particles become polarized, and subsequent dipolar interactions are responsible for aggregating a series of interlinked particles that form chains and columns parallel to the applied field (see Fig. \ref{fig:ERchains}). In the general case \cite{halseyPhysRevLett90,blockLANGMUIR90,halseySCIENCE92,halseySciAmer93,parthasarathyMatSciEng96,weijiaJApplPhys99,HaoAdvMater2001,haoAdvColIntSci2002}, i.e. whatever the nature of the suspensed particles, this kind of structuring stabilizes within some tens of seconds, and disappears when the field is removed. The structuring coincides with a drastic change in the rheological properties of the suspensions, which is why they are called Electro-Rheological (ER) fluids \cite{halseySCIENCE92}; the mechanical behavior of such a system is thus controllable by the applied electric field \cite{halseyPhysRevLett90,blockLANGMUIR90,halseySCIENCE92,halseySciAmer93,parthasarathyMatSciEng96,weijiaJApplPhys99,HaoAdvMater2001,haoAdvColIntSci2002}. Various physical and chemical properties of particles and the suspending liquid may control the behavior of an ER-fluid: Dielectric constant and/or- conductivity of the suspended particles \cite{blockLANGMUIR90,parthasarathyMatSciEng96}, volume fraction of particles \cite{tianMaterLett2003}, frequency and magnitude of the applied electric field \cite{blockLANGMUIR90,parthasarathyMatSciEng96}. Other factors such as particle geometry \cite{KanuJRheol98,DuanJIntellMatSystemsStructures2000,tianMaterLett2003}, size \cite{shihIntJModPhys1994,TanPhysRevE1999} and polydispersity \cite{shihIntJModPhys1994,TanPhysRevE1999} also influence the ER-shear stress behavior. Small amounts of additives like water \cite{wongCONF89}, adsorbed and/or- absorbed on the particles, may also play an important role for certain types of ER fluids. \begin{figure} \centering \includegraphics[width=0.45\textwidth]{Fig1a_tweaked.eps} \caption{\label{fig:ERchains} Microscopy picture of ER chains obtained from applying a $1$kV/mm electric field to a suspension of NaFh clay crystallite aggregates in a silicon oil. The red bands seen in the upper and lower parts of the image are copper electrodes.} \end{figure} Recently, we have studied the effect of an external DC electric field ($\sim 1$ kV/mm) on the rheological properties of colloidal suspensions consisting of laponite particles in a silicone oil \cite{parmarLangmuir2008}. The laponite particle aggregates were observed to exhibit the polarization and subsequent self-assembly typical of ER fluids. Without an applied electric field, the steady state shear behavior of such laponite suspensions is Newtonian-like. Under application of an electric field larger than the triggering field, the samples' rheological properties changes dramatically: a significant yield stress was measured, and under continuous shear the fluid exhibited a shear-thinning behavior. We studied the rheological properties, in particular the dynamic and static shear stress, $\tau$, as a function of particle volume fraction $\Phi$, for various strengths of the applied electric field $E$. We showed that the flow curves under continuous shearing can be scaled with respect to both $\Phi$ and $E$, onto a master curve in the form: \begin{equation} \label{eq:scaling_law} \tau (\Phi, E) = \Phi^\beta E^\alpha f\left ( \frac{\dot{\gamma}}{\Phi^{2/3}\, E^2} \right ) \text{~ ,} \end{equation} where $\dot{\gamma}$ is the shear rate, $\alpha$ and $\beta$ are scaling exponents, and $f$ represents the master curve. We demonstrated that Eq.~\eqref{eq:scaling_law} is consistent with simple scaling arguments. We showed that in the laponite case, the shape of this master curve approaches a standard power-law model at high Manson numbers. For our laponite suspensions, both dynamic and static yield stress were shown to depend on the particle volume fraction and applied electric field as $\Phi^\beta \, E^\alpha$, with $\alpha \sim 1.85$, and $\beta \sim 1.00$ and $1.70$, for the dynamic and static yield stresses respectively. The clay particles investigated in the present study belong to the same family as the laponite particles, namely 2:1 clays, whose basic structural unit is a $1$ nm-thick platelet consisting of two tetrahedral silica sheets sandwiching one octahedral silica sheet. 2:1 clays carry a moderate negative surface charge on their plane surfaces. This charge is usually sufficiently large so that individual platelets be able to stack by sharing cations, and at the same time moderate enough so as to allow further intercalation of water molecules into the resulting “decks of cards”-like smectite particles. Natural 2:1 clay particles dispersed in salt solutions have been studied for decades \cite{veldeBOOK92}, and recently there has been a growing activity in the study of complex physical phenomena in synthetic smectites \cite{fossumPHYSICA99}. Much effort has gone into relating the lamellar microstructure of smectite clay-salt water suspensions to their collective interaction and to resulting macroscopic physical properties, such as phase behavior and rheological properties \cite{GabrielJPhysChem96,mourchidLangmuir98,bonnLangmuir99,DiMasiPRE2001,bonnPRL2002,coussotPRL2002,lemaireEurophysLett2002,fossumENERGY2005,fonsecaJApplCryst2007}. Nematic liquid crystalline-like ordering in smectite systems has been characterized by the observation of birefringent domains with defect textures \cite{GabrielJPhysChem96,lemaireEurophysLett2002,fossumENERGY2005}, by X-Ray scattering \cite{DiMasiPRE2001,fonsecaJApplCryst2007,fonsecaPRE2009,hemmenLangmuir2009} and by NMR spectrometry and imaging \cite{deAzevedoLangmuir2007,hemmenLangmuir2009}. The synthetic 2:1 clay fluorohectorite (FH) is a hectorite clay in which the hydroxyl groups have been replaced by fluorin atoms. Each individual clay platelet is about $1$ nm thick, and its density is $2.8$ \cite{knudsenJApplCryst2003}. Fluorohectorite particles suspended in saline solutions consist of $\sim 80$\textendash $100$ negatively surface-charged clay platelets \cite{kaviratnaJPhysChemSol96,DiMasiPRE2001,dasilvaPRE2002,daSilvaPRB2003} residing on top of each other with intercalated charge-balancing cations X \cite{kaviratnaJPhysChemSol96,fossumPHYSICA99}, such as Na$^+$, Ni$^{2+}$, or Fe$^{3+}$. They have a large aspect ratio: their thickness is $\sim 0.1$ \textmu{}m, while their lateral size can vary from tens of nm up to a few \textmu{}m. This deck-of-card structure of the X-FH clay particles subsists in a weakly-hydrated state. The layer charge density of the individual platelet is 1.2 e$^-/$unit cell \cite{kaviratnaJPhysChemSol96}, which allows the interlayer cations to be exchanged with other types of cations in solution. A fluorohectorite system in which all intercalated cations have been exchanged so as to be of one type X only is denoted X-FH. In the present study, we have used Na-FH, whose half unit cell has a formula Na$_{0.6}$(Mg$_{2.4}$Li$_{0.6}$)Si$_4$O$_{10}$F$_2$. As for all 2:1 clays, water molecules or other polar molecules can be adsorbed or intercalated between interlayer spaces of the platelets, depending upon the environmental conditions such as temperature and relative humidity \cite{kaviratnaJPhysChemSol96,dasilvaPRE2002,daSilvaPRB2003,knudsenJApplCryst2003,knudsenPHYSICA2004,lovollPhysicaB2005,meheustClayScience2005}. When observed in a scanning electron microscope (SEM), untreated sodium fluorohectorite (Na-FH) clay powder display particles with edges and surfaces that are very rough and they are highly anisotropic in their shape, and highly polydisperse in size along their lateral directions. The lateral size of such clay particles can be further reduced by fine grinding of its powder. Recently, we demonstrated that X-FH particles suspended in silicone oil display ER structuring \cite{fossumEPL2006}. We investigated the chain or-column like-structures of X-FH particles \cite{fossumEPL2006} using synchrotron X-ray scattering. We determined the structural morphology (preferred orientations of the X-FH particles) of the chain- or column-like structures and the direction of the induced dipoles in X-FH particles. We pointed to the possible migrations of surface charges and possible movement of interlayer cations X of X-FH as main factors responsible for their polarization in an external DC electric field. In the present report we complement the previous structural studies with results from rheometry measurements under steady shear, following the method previously used in our study of electro-rheological laponite suspensions \cite{parmarLangmuir2008}. \section{Physical background -- The rheology of ER fluids} \subsection{Rheology of ER suspensions when no external electric field is applied} Without an electric field, the ideal steady state shear behavior of an ER fluid is Newtonian-like \cite{KlingenbergLangmuir1990} and an increase of its particle volume fraction increases the magnitude of its viscosity $\eta$. The $\Phi$-dependent viscosity of an ER fluid of monodisperse spherical particles suspended in Newtonian-liquid ($\Phi <10$\%) can be approximated by the Batchelor relation and the Krieger-Dougherty relation \cite{macoskoBOOK94}. These relations account for the interactions between the particles themselves and between particles and the surrounding liquid. One empirical equation which has been found to account for the viscosity of monodisperse and polydisperse sphere suspensions in a range of particle fraction up to $50$\%, has been suggested by Chong, Christiansen and Baer \cite{chongJApplPolymSci71}. The Batchelor and Krieger-Dougherty relations can strictly only be used to describe the viscosity of concentrated colloidal suspensions of isotropic and monodisperse non-charged particles, but the Chong-Christiansen-Baer relation \cite{chongJApplPolymSci71} has been found to be an appropriate description for concentrated suspensions of kaolinite clay \cite{coussotPRL1995}, a system that consists of anisotropic and polydisperse particles. In our recent study of oil suspensions of synthetic laponite we found that the following generalized Krieger-Dougherty relation can be used to fit the zero field viscosity $\eta$ \cite{parmarLangmuir2008} up to $\Phi = 50$ \%: \begin{equation} \label{eq:viscosity_model_by_chong_etal} \eta = \eta_0 \: \left ( 1 - \frac{\Phi}{\Phi_\text{max}}\right )^{-\eta_\text{I}\ \Phi_\text{max}} \text{~ ,} \end{equation} in which the parameter $\eta_\text{I} = \lim_{\Phi \rightarrow 0} \left [ (\eta/\eta_0 -1)/\Phi \right ]$ is the fluid's intrinsic viscosity, and $\eta_0$ is the viscosity of the carrier fluid, while $\Phi_\text{max}$ is the packing value for the volume fraction, which cannot be exceeded. As $\Phi$ increases, the various interaction forces such as hydrodynamics forces, dispersion forces, electrostatic forces, and steric effects/polymeric forces also increase. The net result is an increase in the viscosity of the suspension. \subsection{Rheology when an external electric field is applied} Under application of an electric field ($\sim 1$ kV/mm) and under steady state shear, a general ER fluid shows a well-defined yield stress beyond which it tends to be shear-thinning (pseudoplastic) i.e. the viscosity decreases with increasing shear rate. The typical steady-shear behavior of an ER fluid under the influence of an external electric field is generally characterized as a Bingham-like solid \cite{KlingenbergLangmuir1990}. \additions{This is especially true at large enough shear rates, while at lower shear rates significant deviations are observed, which can be described to some extent by more complex functional relations \cite{zhuJNonNewtonianFluidMech2005,ChoPolymer2005}.} The value of the dynamic yield stress is strongly influenced by the rheological model and the shear rate range selected, as it is obtained by extrapolating the curves from steady state shearing measurement to zero shear rate. This value can be significantly different from the static yield stress i.e. a yield stress of a disrupt (no shearing) ER fluid \cite{shenoyBOOK99}. There are several models that attempt to explain the shear stress behavior for various ER fluids, mostly based on the theoretical analysis of disrupt configurations. The polarization theory \cite{halseyPhysRevLett90,MartinJChemPhys1996}, based on the mismatch in dielectric constant between the particles and the suspending liquid predicts that the yield stress is proportional to $E^\alpha$, where $E$ is the magnitude of the applied electric field and the exponent $\alpha$ is equal to $2$. In the polarization theory, the response of the particle polarization and the suspending liquid are assumed to be linearly related to the applied electric field. The conduction theories based on the mismatch in conductivity between the particles and the suspending liquid predict the exponent as $1\le \alpha \le 2$ at both low- and high- volume fractions of particles \cite{andersonPROCEEDING92,attenIntJModPhys1994,foulcJElectrostat94,WuJPhysD1996,WuIntJModernPhys1996}. Also multipole effects are often proposed to account for particle interactions at relatively high concentrations of particles \cite{blockLANGMUIR90,parthasarathyMatSciEng96}. \subsection{Stress-bifurcation in the rheological behavior} \label{sec:bifurcation_in_viscosity} In our previous study of the electrorheology of laponite suspensions \cite{parmarLangmuir2008}, we have shown that ER suspensions can be studied in the framework of thixotropic fluids. Indeed, their rheology is controlled by a competition between a shear-rejuvenation property and an aging mechanism: the shear-rejuvenation results from the destruction of the ER microstructure, which resists flow-induced particle rearrangement; on the contrary, a suspension left at rest builds up a microstructure in time, which can be considered an aging mechanism, although the particle-particle interactions responsible for it are due to an external electric field. For all thixotropic yield stress fluids, which may be equivalent to saying for all yield stress fluids \cite{mollerSoftMatter2006}, a consequence of this competition mechanism is that different methods of estimating the static yield stress of a fluidlike substance do not necessarily lead to the same result, as it may depend on the shearing history. Coussot and coauthors have proposed a simple phenomenological rheology model \cite{coussotPRL2002,coussotPhysRevLett2002b} accounting for the competition between aging and a shear rejuvenation that is function of the state of the microstructure. In addition to illustrating why the yield stress measured by ``classical'' techniques depends on the history of shear, the model also predicts that there can be shear rates that such fluids cannot accomodate, in other words, that flow localization occurs whenever the fluid is imposed \additions{a given shear rate within the ``forbidden'' range}. Under constant applied shear stress, in contrast, a bifurcation in the flow curves is observed: if the applied stress is smaller than a critical value, the microstructure buildup is dominant, and the viscosity of the fluid eventually goes to infinity; if it is larger, the viscosity goes to a finite value at large times. Shear rate values that are outside of the asymptotic shear rates observed below this bifurcation threshold cannot be accomodated by the fluid without flow localization. This rheology in the bifurcation is actually suggested by Moller et al. \cite{mollerSoftMatter2006} as a principle to estimate the yield stress of a thixotropic fluid in an univoque manner: initally forcing flow in the flow cell and subsequently letting the system evolve under a constant shear stress allows determination of whether this shear stress value is below or above the bifurcation threshold. Investigating various applied shear stress values then leads to an accurate determination of the fluid's yield stress. This technique was successfully applied to a laponite-based ER fluid \cite{parmarLangmuir2008}. \section{Experiments} \subsection{Sample preparation} \label{sec:sample_preparation} As in our previous structural study on X-FH in silicone oil \cite{fossumEPL2006}, we have used synthetic fluorohectorite, originally purchased from Corning Inc. (New York) in powder form and subsequently cation-exchanged to sodium (Na$^+$) \cite{fossumEPL2006}. After dialysis, the Na-fluorohectorite clay was first dried at 100\textcelsius{} for $\sim$10 hours and then grinded in powder form with the help of a mortar and spatula. As suspending liquid in the present studies, we used the silicon oil: Dow Corning 200/100 Fluid (dielectric constant of $2.5$, viscosity of $100$ mPa.s and specific density of $0.973$ at $25$\textcelsius{}) with a small conductivity of about $10$ to $12$ S/m \cite{tangJRheology1995}. Four different ER suspensions of dehydrated Na-FH clay particles with different volume fractions ($\Phi$) were prepared following a protocol previously established with oil-laponite suspensions\cite{parmarLangmuir2008}: (1) removal of water traces in the clay powder and oil by 72 hrs-long heating at $130$\textcelsius{}; (2) mixing of the powders and silicone oil inside glass tubes; (3) cooling of the sealed tubes to room temperature; (4) homogeneization of the suspensions by hand and subsequent ultrasonic bath-shaking for $30$ minutes, at $25$\textcelsius{}. The suspensions were then left to rest so as to remove clay particles with a spherical diameter larger than $\sim 10$ \textmu{}m (according to the Stokes law for particles settling). The remaining suspensions were again hand-shaken about five minutes, three times. It should be noted that after removal of the large particles, the volume fraction $\Phi^\ast$ of the clay particles is significantly lower than the initial volume fraction, $\Phi$, of the prepared ER suspensions. Consequently, $\Phi^\ast$ is not precisely known for the measurements. \additions{The clay powder used in the preparation of the suspensions has been observed to be highly polydisperse. Indeed, TEM imaging of randomly-selected particles from the powders, provided a statistics for the typical particle dimension, defined as the square root of the measured basal area of $151$ imaged particles. We obtained surprisingly well-defined power law-shaped cumulative distributions. We think that the statistics is sufficient to clearly state that the size probability density function in the powder is a monotonically-decreasing function, hence, that the smaller the particles, the more numerous they are. This sort of size distribution is totally different from that of powders from which most ER systems are made, and for which the particle size is selected in a given range and the size probability density function consequently displays a peaked behavior around the average particle size. On the other hand, the distribution of particle thicknesses (perpendicular to their basal sheet), which can be easily inferred from X-ray diffraction measurements \cite{dasilvaPRE2002}, has previously been found to be peaked around $100$ nm. } \additions{ Note however, that once suspended in oil, the clay crystallites tend to aggregate with each other, because their surfaces have more affinity with each other than with the oil. Hence, the ``clay particles'' as they appear in the oil are aggregates of the original clay crystallites. It is difficult to measure the aggregates' dimensions in situ in the oil, as most nano-imaging techniques (electron-microscopy, AFM) cannot easily be used for objects placed in a viscous liquid environment. These aggregates then evolve by deformation and further aggregation after the electric field has been turned on; they then start being visible under the optical microscope, and even by the naked eye. } \subsection{Rheology measurement} \label{sec:rheology_measurements} The rheology of fluorohectorite clay suspensions was measured under an applied DC electric field, using a Physica MCR 300 Rotational Rheometer equipped with a coaxial cylinder electrorheological Couette cell (Physica ERD CC/27). The outer cylinder diameter is $14.46$ mm, and the inner is $13.33$ mm. The immersion length is 40 mm and the sample volume is $19.35$ ml. Note that the present protocol of sample preparation (see section \ref{sec:sample_preparation}) requires relatively large amounts of clay powder in order to prepare this volume of oil suspension. All rheological measurements were carried out at constant temperature $25$\textcelsius{}. Two grounding brushes connected to the bob's axis induce an artificial $1$ Pa yield stress in all data, but this value is negligible compared to all yield stress values addressed here. The steady shear rheological properties were measured after the suspensions had been pre-sheared at $\dot{\gamma} = 200$~s$^{-1}$ for $60$ s in order to ensure the same initial conditions. In order to determine the static yield stress, a control shear stress (CSS) \additions{experiment} was performed: the applied shear stress was increased linearly (by steps of $2$ Pa) on disrupt ER suspensions that had been in applied the DC electric field for 300 s prior to applying the shear stress. We also studied the suspensions in the framework of thixotropic yield stress fluids, measuring the viscosity of samples that were initially forced to flow and then applied a constant shear stress. We have presented the principle of the method in section \ref{sec:bifurcation_in_viscosity}. Due to the competition between shear-induced rejuvenation and the electrically-induced formation of electro-rheological chains, the observed asymptotic rheology can fall into either one rheological behavior or another, depending on the magnitude of the applied shear stress. The bifurcation shear stress that separates those two rheological behaviors is taken as the yield stress of the suspension, and defined pratically as the “stress below which no permanent flow occurs”. \section{Results} \subsection{Rheology of the suspensions in the absence of external electric field} \begin{figure} \centering \includegraphics[width=0.45\textwidth]{viscosity_vs_Phi_nofield-2.eps} \caption{\label{fig:figure1} Viscosity of the Na-FH suspensions as a function of the volume fraction $\Phi$, at $E \sim 0$. The viscosity values have been obtained from linear fits (shown as plain lines) onto the flow curves (shown as symbols) in the inset.} \end{figure} Steady state flow curves (i.e. obtained from Controlled Shear Rate tests) for four different volume fractions ($\Phi$) of Na-FH clay suspensions at zero applied electric field are shown in the inset to Fig.~\ref{fig:figure1}. These suspensions exhibit a Newtonian behavior i.e. a constant viscosity (ratio of the shear stress to the applied shear rate) and no yield stress. The viscosity of the Na-FH clay suspensions increases with the volume fraction of clay particles as illustrated in Fig.~\ref{fig:figure1}. Using Eq.~\eqref{eq:viscosity_model_by_chong_etal}, we find $\eta_I = 0.30$ and $\Phi_\text{m}=61.84$\% as a best fit to the data in Fig.\ref{fig:figure1}. $\Phi_\text{m}$ is expected to be a measure of the maximum packing fraction \cite{chongJApplPolymSci71}. As discussed above in section \ref{sec:sample_preparation}, the true $\Phi^\ast < \Phi$ is uncertain in the present case, thus the meaning of $\Phi_\text{m}$ is somewhat lost (the true maximum packing fraction probably being smaller than $\Phi_\text{m}$ in the same manner as $\Phi^\ast$ is overestimated by $\Phi$). \subsection{CSR tests} \begin{figure} \centering (a) \includegraphics[width=0.425\textwidth]{fig2a.eps} \vfill (b) \includegraphics[width=0.45\textwidth]{fig2b.eps} \\ \caption{\label{fig:figure2}(a) Flow curves of Na-FH suspensions with $\Phi = 37.2$\%, for different strengths of the applied electric fields. (b) Corresponding evolution of the viscosity versus the shear rate.} \end{figure} Fig.~\ref{fig:figure2}(a) and \ref{fig:figure2}(b) show the shear stress- and viscosity- curves (obtained from Controlled Shear Rate CSR tests) for a Na-FH clay suspension of volume fraction $\Phi=37.2$~\%, for various magnitudes of the applied DC electric field. Behaviors similar to Fig.~\ref{fig:figure2}(a) and \ref{fig:figure2}(b) are observed for other prepared concentrations of Na-fluorohectorites particles. Increasing the electric field causes an increase in the shear stress (and viscosity) of these suspensions, while the rheology becomes shear thinning over the entire range of shear rates. The flow curves clearly show the yielding behavior of ER suspensions under application of an electric field: the dynamic yield stress $\tau_0^\text{d}$, defined as the limit towards very low shear rates of the measured yield stress, becomes larger as the magnitude of the electric field increases. \begin{figure} \centering \includegraphics[width=0.45\textwidth]{flow_curves_variousE_FinalScaling-2.eps} \caption{\label{fig:figure3} Scaling of the data of Fig.~\ref{fig:figure2}(a) using Eq.~\eqref{eq:scaling_law} with $\alpha = 1.93$, for a particle fraction $\Phi = 37.2$\%. The plain line represents the asymptotic Bingham model. $E_r=1$ kV is the reference electric field value.} \end{figure} Following the approach previously used for laponite ER- effects under steady shear \cite{parmarLangmuir2008}, we show in Fig.~\ref{fig:figure3} a scaling consistent with Eq.~\eqref{eq:scaling_law} (for a given $\Phi$) applied to the data in Fig.~\ref{fig:figure2}(a). The result is in good agreement with our previous findings for laponite \cite{parmarLangmuir2008}. Fig.~\ref{fig:figure3} corresponds to one concentration $\Phi = 37.2$ \%. The other concentrations that we have investigated display equivalent behaviors with $\alpha \simeq 1.93$. The asymptotic behavior at the highest values of the scaled shear rate, $\dot{\gamma}/E^2$, is Bingham like. \begin{figure} \centering (a) \includegraphics[width=0.39\textwidth]{fig4a-2.eps} \vfill (b) \includegraphics[width=0.40\textwidth]{fig4b-2.eps} \caption{\label{fig:figure4}(a) Flow- curves of Na-FH suspension for two different $\Phi$-values at $E\sim 1.77$ kV/mm. (b) Viscosity- curves for the same suspensions.} \end{figure} Fig.~\ref{fig:figure4} shows the effect of the volume fractions ($\Phi$) of clay particles on the flow- and viscosity- curves (Controlled Shear Rate CSR tests) for Na-FH suspensions at a fixed applied electric field $E \sim 1.77$ kV/mm. The shear stress and viscosity increase with the volume fraction ($\Phi$) of clay particles. In the present case, we have not been able to find a good data collapse and scale the data in Fig.~\ref{fig:figure4} (a) with respect to particle fractions, according to Eq.~\eqref{eq:scaling_law}. \subsection{CSS tests} \begin{figure} \centering \includegraphics[width=0.45\textwidth]{fig5.eps} \caption{\label{fig:figure5} Log-log plot of the static yield stress, $\tau_0^\text{s}$, plotted versus the strength of the applied DC electric field for different volume fractions of the Na-FH particles.} \end{figure} Figure \ref{fig:figure5} shows the static yield stress (obtained from Controlled Shear Stress CSS test) plotted as a function of the applied DC electric field $E$, for different volume fractions ($\Phi$) of the present ER fluids. The static yield stresses, $\tau_0^\text{s}$, increase when the $E$ is increased: all measured static yield stresses are observed to be proportional to $E^\alpha$, with values of exponents $\alpha$ close to $1.58$ for all $\Phi$ values. In Fig.~\ref{fig:figure6}, we have replotted the yield stress, scaled by $(E/E_r)^{1.58}$, where $E_r=1$ kV is a reference electric field value, as a function of the particle volume fraction. We do obtain a rather satisfying data collapse. For particle fractions up to $37.2$\%, the dependence of the scaled data versus $\Phi$ is a power law with an exponent $0.53$. Between $\Phi=37.2$\% and $\Phi=47.0$, a large “jump” in static yield stress is clearly seen; it is experimentally reproducible for the one sample we have studied at a volume fraction $\Phi = 47.0$\%. At present we cannot explain this “jump”, although we can suggest that it be due to formation of larger effective aggregates above some critical threshold $\Phi$ between $37.2$\% and $47.0$\%. \begin{figure} \centering \includegraphics[width=0.45\textwidth]{static_thaus_FH_vsPhi_scaled.eps} \caption{\label{fig:figure6} Log-log plot of the scaled static yield stress versus particle volume fractio . The dashed line (power law with an exponent $1.70$) represents the behavior of our previously-reported laponite suspensions \cite{parmarLangmuir2008}.} \end{figure} \subsection{Bifurcation in the rheological behavior} \begin{figure} \centering (a) \includegraphics[width=0.4\textwidth]{fig7a.eps} \vfill (b) \includegraphics[width=0.4\textwidth]{fig7b.eps} \\ \caption{\label{fig:bifurcationNaFH} Bifurcation in the rheology of a NaFH suspension of volume fraction $\Phi= 33$\%, under applied electric field strengths of $530$ V.mm$^{-1}$ (a) and $885$ V.mm$^{-1}$ (b).} \end{figure} The flow curves obtained when forcing a suspension of volume fraction $\Phi= 33$\% to flow, applying an electric fields of strength $E=530$ V.mm$^{-1}$, and then letting the suspensions evolve under a constant applied stress $\sigma$ are shown in Fig.~\ref{fig:bifurcationNaFH}(a), for various magnitudes of the applied stress. The bifurcation in the rheological behavior occurs for $\tau=\tau_b$ between $20$ and $22$ Pa. The same type of plots are shown for $E=885$ V.mm$^{-1}$ in Fig.~\ref{fig:bifurcationNaFH}(b), leading to $\tau_b$ between $27$ and $29$ Pa. In contrast to our previous study on laponite particles \cite{parmarLangmuir2008}, the method could not be used for larger strength of the applied electric field because the voltage supply could not maintain the required voltage difference throughout the cell due to leak currents running through the cell. However we were able to infer a bifurcation yield stress for 11 different $(\Phi,E)$ configurations (see table~\ref{tab:results_bifurcation_NaFH}). These measured bifurcation stresses are plotted as a function of the electric field in Fig.~\ref{fig:bifurcationSigma_vs_E}. A power law behavior is observed, but in contrast to what was observed for the CSS and CSR tests, the exponent seem to decrease with increasing volume fraction. We believe that this is due leak currents and to the resulting discrepancy between the voltage set on the supply and that really applied to the cell. Indeed, this discrepancy is expected to be all the more important as the volume fraction $\Phi$ becomes larger, and therefore a plot of the yield stress versus the nominal electric field will display an exponent all the smaller (with respect to the one that would be observed on a plot versus a measured voltage) as $\Phi$ is larger. Our Physica rheometer is not equipped to measure leak currents that run across the electrorheological cell of the rheometer. \begin{table} \caption{\label{tab:results_bifurcation_NaFH} Critical stresses measured for various electric field strengths and particle fractions. Leak currents prevent accurate measurement for the two larger electric fields and particle fractions.} \centering \begin{tabular}{c | c c c} \hline \hline $E$ (V.mm$^{-1}$) & $\Phi = 10$\% & $\Phi = 23$\% & $\Phi = 33$\% \\ \hline \hline 532 & 3 & 8 & 21 \\ 708 & 4 & 10 & 23 \\ 885 & 6 & 14 & 28 \\ 1330 & 12 & -- & -- \\ 1720 & 17 & -- & -- \\ \hline \hline \end{tabular} \end{table} \subsection{Discussion} \begin{figure} \centering \includegraphics[width=0.40\textwidth]{fig8-2.eps} \caption{\label{fig:bifurcationSigma_vs_E} Bifurcation yield stress $\tau_b$ as a function of the applied electric field, for three different particle volume fractions. The fitted power law exponents are $1.53$, $1.07$ and $0.54$, respectively. This decrease of the exponent with volume fraction is attributed to the growing discrepancy between the measured- and applied- voltage.} \end{figure} \subsubsection{The electrorheology of Na-FH suspensions} The three types of measurements performed in this study have shown a power-law dependence versus applied electric field of the yield stress measured on a given suspension. For reasons summarized in section \ref{sec:bifurcation_in_viscosity} and presented in more details in the introduction to our previous work on electrorheological laponite suspensions \cite{parmarLangmuir2008}, we believe that the best-posed quantity to experimentally characterize the yielding behavior of a thixotropic fluid is the bifurcation yield stress $\tau_b$ such as defined by Møller et al.\cite{mollerSoftMatter2006}. As previously done for laponite suspensions, we have been able to exhibit the rheological bifurcation of the fluorohectorite suspensions. This is a confirmation that such rheological fluids, independently of their composition, can be studied in the framework of thixotropic rheological models. Besides, we could determine a scaling law in the form $\tau_b \propto E^\alpha$ for all studied suspensions, with $\alpha$ exponents that decrease with the particle fraction. In spite of being well-posed theoretically, the bifurcation method is difficult to perform experimentally because it involves lengthy measurements, allowing phenomena such as sedimentation to occur. In this case, we believe that leak currents are the main reasons for the exponent $\beta$ being dependent on $\Phi$ in Fig.~\ref{fig:bifurcationSigma_vs_E}, as the electric field felt by the particles is weakened with respect to the field applied to the ER cell, and this all the more as the average inter-particle distance is small. This interpretation is consistent with the value $\alpha=1.58$ measured for the static yield stress $\tau_0^\text{s}$ in the range $1.3<E<2.7$ kV (Fig.~\ref{fig:figure5}), which is slightly larger than the $\alpha = 1.53$ measured for $\tau_b$ at $\Phi=10.0$\%. The static yield stress classicly measured with disruption (CSS) tests seems to be consistent with the bifurcation yield stress here. Besides, CSS tests are little affected by sedimentation because they involve a measurement over a short time range, starting with a situation in which all particles are ``frozen'' by the electric field. For particle fractions in the range $19.1\%\leq\Phi\leq37.2\%$, they provide a power-law dependence of $\tau_0^\text{s} \propto \Phi^\beta$, with $\beta=0.54$ (Fig.~\ref{fig:figure6}). From the bifurcation tests and CSR tests, we infer a scaling of the suspension yielding behavior in the form \begin{equation} \tau \propto E^\alpha \, \Phi ^\beta \text{~,} \end{equation} with $\alpha = 1.55 \pm 0.50$ and $\beta = 0.54\pm 0.05$. When confronted to existing models, the value for $\alpha$ suggests that interaction forces between polarized clay particles are governed by their conductivity mismatch to that of the silicone oil, thus indicating that the polarization of such clay particles in an applied external DC electric field is partly caused by migration of surface charges. In a previous study, we had already shown that the movement of interlayer cations (Na$^+$) sandwiched inside the nano-stacked fluorohectorite particles also plays a role in the polarization \cite{fossumEPL2006}. In the present study, the conductivity of dehydrated Na-FH clay particles that are used to prepare the ER suspensions is not known. Note that the conductivity of the suspending oil might also increase under application of an external electric field \cite{tangJRheology1995,foulcJElectrostat94}, and this may lead to a significant effect on the ER behavior. The CSR tests, on the contrary, provide a dynamic yield stress, $\tau_0^\text{d}$, as the extrapolation of the flow curves at low shear rates. It scales with respect to the applied field as a power law with an exponent $\alpha$ closer to $2$ than that obtained from the static and bifurcation yield stresses. This is consistent with previous observation on data from laponite suspensions. Though the meaning of the dynamic yield stress in terms of yielding behavior of the suspensions is not clear, it is remarkable that, as for the laponite ER suspensions, we be able to scale the flow curves corresponding to a given particle fraction $\Phi$ so as to position them on the same master curve, following a scaling law in the form \begin{equation} \tau_d(\Phi,E) = E^\alpha f\left ( \frac{\dot{\gamma}}{E^2} \right ) \text{ ~.} \end{equation} In contrast to the CSR flow curves from laponite suspensions, though, the CSR flow curves from Na-FH suspension could not be scaled as a function of particle volume fractions (following Eq.\eqref{eq:scaling_law}). We attribute this to the large uncertainties connected to the real $\Phi$-values (sedimentation). \subsubsection{ER effect of Na-fluorohectorite suspensions vs. ER effect of laponite suspensions} Though (i) NaFH crystallites have a larger structural surface charge density ($1.2$e$^-/$unit cell) than laponite crystallites ($0.4$e$^-/$unit cell), and (ii) the anisotropic shape of NaFH particles may cause enhanced dipolar interactions as compared to isotropic spherical particles of same size and same conductivity \cite{KanuJRheol98}, the measurements presented in the present study show that the ER effect results in a weaker yield stress fluid in the case of fluorohectorite particles than what was observed with laponite particles. We believe that the size and shape polydispersity of the clay particle population potentially makes the ER aggregates weaker, leading therefore to a reduction in shear stress- and viscosity- as these will depend on the packing of individual clay particles in chain or- column like-structures \cite{brooksColloidsSurf1986}. \begin{table} \centering \caption{\label{tab:exponent_comparison} Exponent $\alpha$ and $\beta$, as defined by Eq.~\eqref{eq:scaling_law}), obtained for ER laponite suspensions (results taken from ref.\cite{parmarLangmuir2008}) and NaFH ER suspensions (this study) from our three yield stress estimates: dynamic-, static-, and bifurcation- yield stress.} \begin{tabular}{l || c | c || c | c} \hline \hline & \multicolumn{2}{c||}{Laponite} & \multicolumn{2}{c}{Na-fluorohectorite} \\ \hline & $\alpha$ & $\beta$ & $\alpha$ & $\beta$ \\ \hline \hline Dynamic y. s. & 1.85 & 1.00 & 1.93 & --\\ Static y. s. & 1.85 & 1.70 & 1.58 & 0.54\\ Bifurcation y. s. & 1.84 & 1.70 & \textit{0.5-1.6} & --\\ \hline \hline \end{tabular} \end{table} In terms of exponents of the scaling law, the $\alpha$ values are similar for both systems, for all estimates of the yield stress (see table~\ref{tab:exponent_comparison}). The $\beta$ values for fluorohectorite suspensions with $\Phi \leq 37.2$, on the contrary, are significantly lower than what was observed for laponite suspensions. In the latter ER fluids, the ``particles'' consist of aggregates of $1$nm-thick laponite platelets. From our prevous study \cite{parmarLangmuir2008}, we concluded that they are three-dimensional in nature. Here, prior to application of the electric field, the fluorohectorite particles are also aggregates of clay crystallites that are quasi-two-dimensional (being nano-stacks of average aspect ratio 10:1), but the crystallite population is very polydisperse, which probably leads to much larger shape and size variations in the clay aggregates in the oil. It is possible that at sufficiently large volume fraction, the aggregates become quasi-spherical in a way similar to those for laponite. This could explain the change of behavior observed for the largest particle fraction in Fig.~\ref{fig:figure6}. \section{\label{sec:Conclusion and prospects} Conclusion and prospects} We have studied the electrorheological behavior of ER fluids consisting of sodium fluorohectorite clay particles suspended in silicone oil. In the absence of electric field, these suspensions are Newtonian-like, and their viscosity is controlled by the particle volume fraction. Under application of a DC electric field larger than about 0.4 kV/mm, the flow curves are Bingham-like, exhibiting a well-defined yield stress that increases with increasing volume fraction of particles as well as with increasing applied electric field. The static and bifurcation yield stresses observed follow a power law $E^\alpha$, with $\alpha\sim1.5 - 1.6$ suggesting that interaction forces are governed by the conductivity mismatch between dipolar particles (fluorohectorite) and the suspending medium (silicone oil). The observed yield stresses are comparable to similar systems (see table 1 in ref.~\cite{parmarLangmuir2008}). This works complements our previous study on these systems \cite{fossumEPL2006} with mechanical behaviors. The presently-obtained values for yield stresses suggest that such fluorohectorite-based ER fluids would not be able to compete with other systems in terms of the practical uses of ER fluids. Rather, one application of these systems of (possibly functionalized) fluorohectorite particles could be in guided self-assemblies of nanoparticles for nano-templating and/or inclusion in composite materials. In order to discriminate the effects of collective properties (geometry and cohesion of the packings) from that of individual particle properties (physical shape and chemical polarisibility) on the suspension's electrorheology, controlling the polydispersity of the suspensions and systematically studying samples based on several types of clay minerals is necessary, which also implies being able to measure aggregate size distributions in situ, in the oil. Such a systematic investigation is beyond the present study. Another prospect is the functionalization of the clay surfaces in order to ease single grain/nano-stack dispersion in oils. \section*{Acknowledgments} This work was supported by NTNU, and by the Research Council of Norway (RCN) through the RCN NANOMAT Program) as well through a RCN SUP (Strategic University Program) project awarded to the COMPLEX Collaborative Research Team in Norway. Y. M. acknowledges the Egide organization for financial support in traveling between France and Norway, under the Aurora framework (grant nr. 18810WC). \additions{J. O. F. acknowledges travel support from the RCN under the same Aurora framework.} \bibliographystyle{elsarticle-harv}
\section{Introduction} One of the reasons for the growing interest in granular materials, i.e. collections of interacting macroscopic particles \cite{aronson,pouliquen,silbert,lutsko01,alam03,GDRMiDi,lutsko04,namiko05,kumaran,orpe07,hisao07a,namiko07,hisao07b,hayakawa08,kudrolli,Hatano08,Kumaran09,Hayakawa09s,Otsuki09s,Otsuki09r,Otsuki09,Otsuki:Jstat,Otsuki:EPJE,Luding09} is the fact that these materials are different from ordinary matter \cite{jaeger}. The pertinent differences do not preclude a description of (up to) moderately dense and nearly elastic granular flows by hydrodynamic equations with constitutive relations derived using kinetic theory \cite{lutsko04,hisao07a,Brilliantov,JR85a, Dufty, Santos, Garzo, Soto, Brey, Goldhirsch03, Lutsko05}. When nontrivial correlations, such as long-time tails and long-range correlations, are present, one can apply fluctuating hydrodynamic descriptions to granular fluids and the latter can be obtained from kinetic theory as well. \cite{kumaran,hisao07b,kudrolli,Kumaran09,Hayakawa09s,Otsuki09s,Otsuki09r,Otsuki09,Otsuki:Jstat,Otsuki:EPJE}. Similar analysis cannot be applied to systems near the jamming transition. Indeed, we know many examples when the behavior of very dense flows cannot be understood by Boltzamnn-Enskog theory \cite{Ishiwata,Garcia,namiko07,Khain07,Khain09,Luding09,Otsuki:PTP, Otsuki:PRE} due to effects like ordering or crystallization, excluded volume, anisotropy and higher order correlations. Therefore, to understand the rheology of dense granular flows, such as the frictional flow \cite{silbert}, and the jamming transition itself \cite{Liu}, an alternative approach is called for. Recently, Otsuki and Hayakawa have proposed a mean-field theory to describe the scaling behavior close to the jamming transition \cite{Otsuki:PTP,Otsuki:PRE} at density (area fraction) $\phi_J$. They predicted that both pressure and viscosity are proportional to $(\phi_J - \phi)^{-4}$. Therefore, the scaled pressure, divided by the kinetic granular temperature ${T} \propto (\phi_J - \phi)^{-2}$, is proportional to $(\phi_J-\phi)^{-2}$, while the scaled viscosity, divided by $\sqrt{T} \propto (\phi_J - \phi)^{-1}$, is proportional to $(\phi_J-\phi)^{-3}$, irrespective of the spatial dimension. The validity of this prediction has been confirmed by extensive molecular dynamics simulations with soft disks. However, one can note that this prediction differs from other results on the divergence of the transport coefficients \cite{Garcia,Otsuki:PTP,Otsuki:PRE,Losert,Olsson}. In particular, Garcia-Rojo et al. \cite{Garcia} concluded that the viscosity for two-dimensional monodisperse rigid-disks is proportional to $(\phi_\eta - \phi)^{-1}$, where $\phi_\eta$ is the area fraction of the 2D order-disorder transition point, while the pressure diverges at a much higher $\phi_P$ with $p \propto (\phi_P-\phi)^{-1}$ \cite{Luding01,Luding01v2,Luding02,Luding04,Luding09}. Not only is the location of the divergence different, but also the power law differs from the mean field prediction in Refs.\ \citen{Otsuki:PTP,Otsuki:PRE}. {\em How can we understand these different predictions?} One of the key points is that the situations considered are different from each other. As stated above, Garcia-Rojo et al. \cite{Garcia,Luding09} used two-dimensional monodisperse rigid-disks without or with very weak dissipation, whereas Otsuki and Hayakawa \cite{Otsuki:PTP,Otsuki:PRE} discussed sheared polydisperse granular particles with a soft-core potential and rather strong dissipation. In order to obtain an unified description on the critical behavior of the viscosity and the pressure in granular rheology, we numerically investigate sheared and weakly inelastic soft disks for both the monodisperse and the polydisperse particle size-distributions. The organization of this paper is as follows: In the next section, we summarize the previous estimates for the pressure and the viscosity for dense two-dimensional disk systems. In Sec.\ \ref{Numerical:Sec}, we present our numerical results for soft inelastic disks under shear in three subsections: In Sec.\ \ref{Model:Subsec}, the numerical model is introduced, Sec.\ \ref{Mono:Subsec} is devoted to results on monodisperse systems, and Sec.\ \ref{Poly:Subsec} to polydisperse systems. In Sec.\ \ref{Time:Subsec}, a criterion for the ranges of validity of the different predictions about the divergence of the viscosity and the pressure is discussed. We will summarize our results and conclude in Sec.\ \ref{Conclusion:Sec}. \section{Pressure and viscosity overview} \label{Previous:Sec} In this section, we briefly summarize previous results on the behavior of pressure and viscosity in two-dimensional disks systems. Following Ref.~\citen{Luding09}, we introduce the non-dimensional pressure \begin{eqnarray} P^* & \equiv & P /(nT) - 1 , \label{P:def} \end{eqnarray} where $P$ is the pressure, $n$ is the number density, and $T=\langle m (\bv{v} - \langle \bv{v} \rangle)^2 \rangle /(2N)$ is the kinetic temperature (twice the fluctuation kinetic energy per particle per degree of freedom) which is proportional to the square of the velocity fluctuations of each particle. We also introduce the non-dimensional viscosity \begin{eqnarray} \eta^* & = & \eta/(\rho v_T s_0/2) \label{eta:def} \end{eqnarray} where $\rho$ denotes the particles' material density, $\rho^B=\rho\phi$ is the bulk area density, the fluctuation velocity is denoted by $v_T = \sqrt{2T/m}$, $s_0=\sqrt{2 \pi}\sigma / 8$, the mass of a grain (we assume all grains to have the same mass) is denoted by $m$, and the mean diameter of a grain (disk) is denoted by $\sigma$. It should be noted that $\rho v_T s_0/2$ is the viscosity for a monodisperse rigid-disk system in the low-density limit and correct to leading order in the Sonine polynomial expansion. For later use, we also introduce the mean free time $t_E$ which is defined as the time interval between successive collisions. This leads to the collision rate $t_E^{-1} = v_T \phi g(\phi) / s_0 = v_T/\lambda$ in the case of dilute and moderately dense systems of rigid disks, where $\lambda$ is proportional to the mean free path. In the first part of this section, let us summarize previous results for elastically interacting rigid disk systems. In the second part of this section, we show other previous results for soft granular disk systems under shear. \subsection{Rigid disk system in the elastic limit} \label{Hard:Subsec} For the equilibrium monodisperse rigid-disk systems, the reduced pressure $P^*$ of elastic systems at moderate densities $\phi < 0.67$ is well described by the classical Enskog theory \cite{Fingerle,Luding01,Luding01v2,Luding04,Luding09} \begin{equation} P^*_4 = 2 \phi g_4(\phi). \label{P4:eq} \end{equation} with the aid of improved pair-correlation function at contact \begin{equation} g_4(\phi) = g_2(\phi) - \frac{\phi^3/16}{8(1-\phi)^4} ~, \label{eq:g4} \end{equation} where $g_2(\phi) = \frac{1-7\phi/16}{(1-\phi)^2}$ in Eq.\ (\ref{eq:g4}) was proposed by Henderson in 1975 \cite{Henderson}. In the regime of high density $\phi > 0.65$, the reduced pressure becomes, first, lower than \eqref{P4:eq} because of ordering (crystallization) and, second, diverges at a density $\phi_P$ due to excluded volume effects. This behavior is quantitatively fitted by \begin{equation} P^*_{\rm dense} = \frac{2 \phi_P }{\phi_P - \phi} h(\phi_P - \phi) -1, \end{equation} with $\phi_P = \pi / (2 \sqrt{3})$, $h(x) = 1 + c_1 x + c_3 x^3$, and the fitting parameters $c_1 = -0.04$, and $c_3 = 3.25$ \cite{Luding09,Luding01,Luding01v2,Herbst}. As shown in references~\citen{Luding09,Luding01,Luding01v2} an interpolation law between the predictions for the low and the high density regions: \begin{equation} P^*_Q = P^*_4+ M(\phi) [P^*_{\rm dense} - P^*_4], \label{P_Q} \end{equation} with $M(\phi) = [1 + \exp ( -(\phi - \phi_c)/m_0]^{-1}$, $\phi_c = 0.699$, and $m_0 = 0.0111$, fits well the numerical data for $P^*$. The quality of the empirical pressure function $P^*_Q$ is perfect, except for the transition region, for which deviations of order of 1\% are observed in the monodisperse, elastically interacting rigid disk system. The dimensionless viscosity for monodisperse elastically colliding rigid disks is well described by the Enskog-Boltzmann equation \begin{equation} \eta^*_E = \left [ \frac{1}{g_2(\phi)}+ 2 \phi + \left (1 + \frac{8}{\pi} \right ) \phi^2 g_2(\phi) \right ]. \label{etaE:eq} \end{equation} Note that $g_2(\phi)$ satisfies $g_2(\phi) \approx g_4(\phi) \approx g_Q(\phi) = P^*_Q / (2 \phi)$, for $ \phi \ll \phi_\eta$. A dominant correction, see Eq.\ (\ref{eta:eq}) below, controls the viscosity for higher densities, closer to $\phi \approx \phi_\eta$. Equation \eqref{etaE:eq} can be used for low and moderate densities, but it is not appropriate close to the crystallization area fraction $\phi_c$ \cite{Ishiwata,Garcia,Khain07,Khain09,Luding09,Otsuki:PTP,Otsuki:PRE}. Therefore, an empirical formula for $\eta^*$ has been proposed as \begin{equation} \eta^*_L = \left (1 + \frac{c_\eta}{\phi_\eta - \phi} - \frac{c_\eta}{\phi_\eta} \right )\eta^*_E, \label{eta:eq} \end{equation} which can fit the numerical data for $0 < \phi < \phi_\eta$ with two fitting parameters $c_\eta = 0.037$ and $\phi_\eta = 0.71$ \cite{Luding09}. Note that the last term is an improvement of the original empirical fit \cite{Garcia} that makes $\eta^*_L$ approach unity for $\phi \rightarrow 0$. Note that $\eta^*$ in Ref.\ \citen{Garcia} was obtained from a non-sheared system by using Einstein-Helfand relation \cite{Helfand}. A slightly different empirical form for the non-dimensional viscosity was proposed by Khain \cite{Khain09} (based on simulations of a sheared system): \begin{equation} \eta^*_K = \left (1 + \frac{c_\eta}{\phi_\eta - \phi} \left ( \frac{\phi}{\phi_\eta} \right )^3 \right ) \eta^*_E ~, \label{etaK:eq} \end{equation} with the same $c_\eta$ and $\phi_\eta$ as before. The reasons for the difference between the viscosity in a sheared and a non-sheared system is an open issue and will not be discussed here. We also introduce the scaled temperature given by \begin{equation} T^* = \frac{T(1-e^2)}{m \dot \gamma^2 s_0^2} \end{equation} for sheared inelastic rigid-disks, where $e$ and $\dot\gamma$ are the coefficient of restitution and shear rate, respectively. Luding observed that the empirical expression \begin{equation} T_K^* = \frac{\eta^*_K}{\phi^2 g_2(\phi)} \label{TK} \end{equation} fits best the numerical data for monodisperse rigid disks \cite{Luding09}, while \begin{equation} T_L^* = \frac{\eta^*_L}{\phi^2 g_2(\phi)} \label{TL} \end{equation} slightly overpredicts the scaled temperature. For polydisperse elastic rigid-disk systems, many empirical expressions for the reduced pressure $P^*$ have been proposed, see e.g.\ \cite{Luding09,Luding01,Luding02,Luding04,Torquato95}. It is known that $P^*$ diverges around $\phi_{\rm max} \simeq 0.85$ for bi- and polydisperse systems, but there is no theory to our knowledge that predicts the dependence of $\phi_{\rm max}$ on the width of the size distribution function that was observed in rigid-disk simulations \cite{Luding01,Luding02}. Dependent on the dynamics (rate of compression), on the material parameters (dissipation and friction), and on the size-distribution, different values of $\phi_{\rm max}$ can be observed. In several studies, the critical behavior was well described asymptotically by a power law \begin{equation} P_d^* \sim (\phi_{\rm max} - \phi)^{-1} \label{Pd:eq} \end{equation} see Refs.\ \citen{Torquato95,Luding01,Luding02}. No good empirical equation for the viscosity of polydisperse rigid-disk systems in the elastic limit has been proposed to our knowledge. However, if we assume that the viscosity behaves like that of the monodisperse rigid-disk system, we can introduce the empirical expression \begin{equation} \eta_{d}^* \sim (\phi_{\rm max} - \phi)^{-1} \label{etad:eq} \end{equation} as a guess. Here, we assume that the pressure $P^*$ and the viscosity $\eta^*$ for the polydisperse system diverge at the same point $\phi_{\rm max}$, which differs from the case of the monodisperse system, where $P^*$ and $\eta^*$ diverge at different points $\phi_P$ and $\phi_\eta$ due to the ordering effect. \subsection{Soft-disk system} \label{Soft:Subsec} Let us consider a sheared system of inelastic soft-disks characterized by the non-linear normal repulsive contact force $k\delta^\Delta$ with power $\Delta$, where $k$ and $\delta$ are the stiffness constant and the compression length (overlap), respectively. For this case, Otsuki and Hayakawa~\cite{Otsuki:PTP,Otsuki:PRE} proposed scaling relations for the kinetic temperature $T$, shear stress $S$, and pressure $P$, near the jamming transition point $\phi_J \simeq 0.85$: \begin{equation} \label{scaling} T = |\Phi|^{x_{\Phi}} {\cal T}_{\pm}\left(\frac{\dot\gamma}{|\Phi|^{\alpha}}\right), \ S = |\Phi|^{y_{\Phi}}{\cal S}_{\pm}\left(\frac{\dot\gamma}{|\Phi|^{\alpha}}\right), \ P = |\Phi|^{y_{\Phi}'}{\cal P}_{\pm}\left(\frac{\dot\gamma}{|\Phi|^{\alpha}}\right), \end{equation} where $\Phi \equiv \phi - \phi_J$ is the density difference from the jamming point. This scaling ansatz is based on the idea that the system has only one relevant time-scale $\tau \sim |\Phi|^{-\alpha}$ diverging near the transition point $\phi_J$, and the behavior of the system is dominated by the ratio of the time scale $\tau$ and the inverse of the shear rate $\dot \gamma$. This idea is often used in the analysis of critical phenomena. The scaling functions ${\cal T}_{+}(x)$, ${\cal S}_{+}(x)$, and ${\cal P}_{+}(x)$ satisfy \begin{eqnarray}\label{jammed} \lim_{x\rightarrow 0}{\cal T}_+(x) & =& x, \quad \lim_{x\rightarrow 0}{\cal S}_{+}(x) = 1, \quad \lim_{x\rightarrow 0}{\cal P}_{+}(x) = 1 \end{eqnarray} for $\phi > \phi_J$, i.e., for higher area fraction. The pressure and shear stress scaling -- in this limit -- represent the existence of a (constant) yield stress $S=S_Y$. The scaling for the temperature is obtained from the assumption that a characteristic frequency, $\omega \equiv \dot \gamma S / (nT)$, is finite when $\dot \gamma \to 0$ in the jammed state $\phi > \phi_J$, see Ref.\ \citen{Wyart}. \footnote{Here, we should note that $\omega$ is proportional to the Enskog collision rate $\omega = (1-e^2) t_E^{-1} /2$, see Ref.\ \citen{Luding09}, in the unjammed state well below the jamming point, $\phi <\phi_J$, i.e., in the collisional flow regime. Due to the prefactor $(1-e^2)/2$, we can identify $\omega$ with the characteristic dissipation rate. The different time-scales (inverse frequencies) and their relative importance are discussed below in subsection \ref{subsubsec:dimensionless}. } On the other hand, for lower area fraction, ${\cal T}_{-}(x)$, ${\cal S}_{-}(x)$, and ${\cal P}_{-}(x)$ satisfy \begin{eqnarray}\label{unjammed} \lim_{x\rightarrow 0}{\cal T}_{-}(x) & = & x^2, {\quad} \lim_{x\rightarrow 0}{\cal S}_{-}(x) = x^2, \quad \lim_{x\rightarrow 0}{\cal P}_{-}(x) = x^2 \end{eqnarray} for $\phi \ll \phi_J$, which represent Bagnold's scaling law in the liquid phase. Furthermore, for diverging argument $x$, i.e., at the jamming point J with $\Phi \rightarrow 0$, the scaling functions ${\cal T}_{\pm}(x)$, ${\cal S}_{\pm}(x)$, and ${\cal P}_{\pm}(x)$ should be independent of $\Phi$ and thus satisfy: \begin{equation}\label{critical} \lim_{x\rightarrow \infty}{\cal T}_{\pm}(x) = x^{x_\Phi/\alpha}, \ \lim_{x\rightarrow \infty}{\cal S}_{\pm}(x) = x^{y_\Phi/\alpha}, \ \lim_{x\rightarrow \infty}{\cal P}_{\pm}(x) = x^{y_\Phi'/\alpha} \,. \end{equation} The critical exponents in Eq.(\ref{scaling}) are given by \begin{equation} \label{exponents} x_{\Phi} = 2+\Delta, \quad y_{\Phi} = y_{\Phi}' = \Delta, \quad {\rm and~} \quad \alpha=\frac{\Delta+4}{2}, \end{equation} which depend on some additional assumptions\cite{Otsuki:PTP}, such as the requirement that the pressure $P$ for $\dot \gamma \to 0$, in the jammed state $\Phi >0$, scales with the force power-law as $P \sim \Phi^{\Delta}$, see Refs.\ \citen{OHern,OHern03}. Thus, the temperature $T$, the shear stress $S$, and the pressure $P$, below the jamming transition point in the zero shear limit $\dot \gamma \to 0$ obey: \begin{equation} \label{scale_T_S_P} T \sim (\phi_J - \phi)^{-2} \dot \gamma ^2, \quad S \sim (\phi_J - \phi)^{-4} \dot \gamma ^2, \quad P \sim (\phi_J - \phi)^{-4} \dot \gamma ^2. \end{equation} Both the viscosity $\eta=S/\dot\gamma$ and pressure $P$, at the jamming transition point, diverge proportional to the area fraction difference to the power $-4$. Substituting Eqs.\ \eqref{scale_T_S_P} into Eqs.\ \eqref{P:def} and \eqref{eta:def}, the reduced pressure $P^*$ and the dimensionless viscosity $\eta^*$, in the vicinity of the jamming point are respectively given by \begin{eqnarray} P_{J}^* & \sim & (\phi_J - \phi)^{-2}, \label{scaling:P}\\ \eta_{J}^* & \sim & (\phi_J - \phi)^{-3}. \label{scaling:eta} \end{eqnarray} It is remarkable that the scaling relations \eqref{scale_T_S_P}--\eqref{scaling:eta} below the jamming transition point are independent of $\Delta$, even though the exponents in Eq.\ \eqref{exponents} depend on $\Delta$. The validity of Eqs.\ \eqref{scaling:P} and \eqref{scaling:eta} for various $\Delta$ has been numerically verified \cite{Otsuki:PTP,Otsuki:PRE}. However, the conjecture that the scaling relations \eqref{scaling:P} and \eqref{scaling:eta} are applicable in the hard disk limit seems to be in conflict with the empirical relations Eqs.\ \eqref{Pd:eq} and \eqref{etad:eq} for elastic rigid-disk systems. \section{Numerical results} \label{Numerical:Sec} In this section, we numerically investigate the reduced pressure $P^*$ and viscosity $\eta^*$ of sheared systems with soft granular particles, with special focus on the rigid-disk limit. In the first part, our soft-disk model is introduced. In the second part, we present numerical results for monodisperse systems, while in the third part the results for polydisperse systems are presented. \subsection{The soft-disk model system} \label{Model:Subsec} \subsubsection{Contact forces and boundary conditions} Let us consider two-dimensional granular assemblies under a uniform shear with shear rate $\dot \gamma$. Throughout this paper, we assume that granular particles are frictionless, without any tangential contact force acting between grains. For the sake of simplicity, we restrict ourselves to the linear contact model with $\Delta=1$. We assume that all particles have identical mass regardless of their diameters. The linear elastic repulsive normal force between the grains $i$ and $j$, located at $\bv{r}_i$ and $\bv{r}_j$, is: \begin{eqnarray} f_{\rm el}(r_{ij}) & = & k \Theta \left( \sigma_{ij} - r_{ij} \right ) (\sigma_{ij}-r_{ij}), \label{elastic:force} \end{eqnarray} where $k$ and $r_{ij}$ are the elastic constant and the distance between the grains $r_{ij}\equiv |\bv{r}_{ij}|=|\bv{r}_i-\bv{r}_j|$, respectively. $\sigma_{ij} = (\sigma_i + \sigma_j)/2$ is the average of the diameters of grains $i$ and $j$. The Heaviside step function $\Theta(x)$ satisfies $\Theta(x) = 1$ for $x \geq 0$ and $\Theta(x) = 0$ otherwise. The viscous contact normal force is assumed as \begin{eqnarray} f_{\rm vis}(r_{ij}, v_{ij,{\rm n}}) & = & - \zeta \Theta \left( \sigma_{ij} - r_{ij} \right ) v_{ij,{\rm n}}, \label{dis:lin} \end{eqnarray} where $\zeta$ is the viscous parameter. Here, $v_{ij,{\rm n}}$ is the relative normal velocity between the contacting grains $v_{ij,{\rm n}}\equiv (\bv{v}_i-\bv{v}_j)\cdot\bv{r}_{ij}/r_{ij}$, where $\bv{v}_i$ and $\bv{v}_j$ are the velocities of the centers of the grains $i$ and $j$, respectively. In order to obtain a uniform velocity gradient $\dot\gamma$ in $y$ direction and macroscopic velocity only in $x$ direction, we adopt the Lees-Edwards boundary conditions. The average velocity $\bv{c}(\bv{r})$ at position $\bv{r}$ is given by $\bv{c}(\bv{r}) = \dot \gamma y \bv{e}_x$, where $e_{x,\alpha}$ is a unit vector component given by $e_{x,\alpha} = \delta_{x\,\alpha}$, where $\alpha$ is the Cartesiani coordinate. \subsubsection{Discussion of dimensionless quantities} \label{subsubsec:dimensionless} There are several non-dimensional parameters in our system. One is the restitution coefficient $e$ given by \begin{equation} e \equiv \exp \left [-\frac{\pi \zeta}{\sqrt{2 k/m - (\zeta/m)^2}} \right ] = \exp \left [ - \zeta t_c \right ] ~, \end{equation} with the pair-collision \footnote{The contact duration $t_c$ is well defined for two masses connected by a linear spring-dashpot system and corresponds to their half-period of oscillation. A particle in a dense packing (connected to several masses by linear spring-dashpots) has a somewhat higher oscillation frequency, but the order of magnitude remains the same. Particles with non-linear contact models can have a pressure dependent $t_c$, but are not considered here. } contact duration $t_c \equiv \pi / \sqrt{2 k/m - (\zeta/m)^2}$. Another is the dimensionless contact duration \begin{equation} \tau_c^* \equiv t_c \dot \gamma \label{tau:def} \end{equation} that represents the ratio of the two ``external'' time-scales of the system \footnote{One can see $\tau^*_c = (\sigma \dot \gamma) / (\sigma /t_c)$ also as the ratio of the two relevant velocities in the dense limit, i.e., as the ratio of the local velocity of horizontal layers that are a diameter of a grain, $\sigma$, apart, and the local information propagation speed $\sigma/t_c$ in a dense packing. However, the ratio of velocities makes only sense in the dense, soft regime, since $t_c$ is not a relevant time-scale in the dilute, near-rigid regime. }. ``External'' means here that these time scales are externally controllable, i.e., the contact-duration is a material parameter and the inverse shear rate is externally adjustable. In all cases studied later, we have $\tau^*_c \ll 1$, which means that the shear time scale is typically much larger than the contact duration, i.e., we do {\em not} consider the case of very soft particles, which is equivalent to extremely high shear rates. Therefore, $\tau^*_c$ will be used as dimensionless control parameter in order to specify the magnitude of stiffness: The rigid disks are reached in the limit $\tau^*_c \to 0$. The third time-scale, $t_E$, in the system is an ``internal'' variable, i.e., cannot be controlled directly. This time scale is proportional to the inverse characteristic frequency of interactions, i.e., the mean free time, $t_E$, in the dilute case or the rigid-disk limit. This defines the (second) dimensionless ratio of times \begin{equation} \tau_E^* \equiv t_E \dot \gamma \label{tauE:def} \end{equation} relevant in the dilute, collisional regime. The third dimensionless number is defined as the ratio of contact duration and mean free time, \begin{equation} \tau_{cE}^* \equiv \frac{t_c}{t_E} = \frac{\tau^*_c}{\tau^*_E} ~. \label{tauc:def} \end{equation} see Eq.\ (53) in Ref.\ \citen{Luding09}. The meaning of this dimensionless number is as follows: For very small $\tau_{cE}^* \ll 1$ one is in the binary collision regime, for large $\tau_{cE}^* > 1$, one is in the solid-like regime with long-lasting multi-particle contacts. In the hard disk limit $\tau_c^*\to 0$, we can identify $\tau_{cE}^*$ with the coordination number as will be shown in Fig.\ \ref{fig:Z}. Namely, finite $\tau_{cE}^*$ in the near-rigid situation means that the system is in a jammed phase. The binary collision regime, $\tau_{cE}^* \rightarrow 0$, cannot be controlled directly, since $t_E$ is a function of temperature, which depends on $e$ and $\dot \gamma$. On the other hand, the rigid-disk limit, $\tau_{c}^* \to 0$, can be approached/realized by either ($i$) vanishing shear rate, $\dot \gamma \to 0$, or ($ii$) near-rigid particles with high stiffness, $k \to \infty$ (with controlling the variable $\zeta$ to maintain a constant restitution coefficient $e$). \begin{table}[htb] \begin{tabular}{|l|l|l|l|c|c|c|c|} \hline & ratio of times & ratio of velocities / stresses & regime of relevance\\ \hline \hline $\tau_c^*$ & $t_c / \dot \gamma^{-1}$ & $v_{\sigma\dot\gamma}/v_c = \sigma \dot \gamma / (\sigma /t_c)$ & near-rigid, high density ($\sigma \gg \lambda$, $t_c \gg t_E$)\\ \hline \hline $\tau_E^*$ & $t_E / \dot \gamma^{-1}$ & $v_{\lambda\dot\gamma}/v_E = \lambda \dot \gamma / (\lambda/t_E)$ & rigid, low density ($\sigma \ll \lambda$, $t_c \ll t_E$)\\ $\tau_{cE}^*$ & $t_c / t_E $ & $v_E/v_c = \sigma/t_E / (\sigma/t_c)$ & near-rigid, low and moderate densities \\ \hline \hline $\tau_\omega^*$ & $\omega^{-1} / \dot \gamma^{-1}$ & $nT/S = 2\tau_E^* / (1-e^2)$ & well defined in sheared systems\\ $\tau_{c\omega}^*$ & $t_c / \omega^{-1} $ & $t_c \dot\gamma S/(nT) = \tau_c^* / \tau_\omega^*$ & well defined in all systems\\ \hline \end{tabular} \caption{ Summary of the dimensionless numbers discussed in the text, where $t_c$, $\dot \gamma^{-1}$, $t_E$, $\omega^{-1}$ are contact duration, inverse shear rate, mean free time, and inverse characteristic dissipation rate, respectively. The velocities $v_{L\dot\gamma}$, $v_c$, and $v_E$ are the shear velocity of layers separated by length $L$, the speed of sound propagation in a dense packing, and the speed of sound propagation in a dilute packing, respectively. The relevant lengths $L$ can be the diameter $\sigma$ (in the dense limit), the mean free path $\lambda=\lambda(\phi)$ (in the dilute limit), or their sum (for all densities). } \label{tab:tab1} \end{table} Furthermore, we can introduce dimensionless numbers that are related to the inverse characteristic dissipation rate $\omega^{-1}$ \footnote{Note that the identity $\omega^{-1} = 2t_E/(1-e^2)$ is true in the dilute, collisional limit only. For higher densities and for softer particles, one has $\omega^{-1} > 2 t_E/(1-e^2)$, i.e., energy dissipation becomes somewhat slower when approaching the jamming transition. This is consistent with a slower energy decay due to the reduced dissipation rate, proposed in Eq.\ (52) in Ref.\ \citen{Luding09}}, which has the meaning of the energy dissipation time-scale. For $e \rightarrow 1$, dissipation is becoming very slow, while for small $e \sim 0$, considerable energy can be dissipated, within a time of order of $t_E$ or $t_c$. Replacing $t_E$ by $\omega^{-1}$ in Eqs.\ (\ref{tauE:def}) and (\ref{tauc:def}), we obtain \begin{eqnarray} \tau_\omega^* & \equiv & \dot \gamma/\omega \,, \label{eq:tauw} \\ \tau_{c\omega}^{*} & \equiv & t_c \omega \,. \label{taucEd:def} \end{eqnarray} It should be noted that $\tau_\omega^*$ and $\tau_{c\omega}^*$ approximately satisfy the relations $\tau_\omega^* \approx 2\tau_E/(1-e^2)$ and $\tau_{c\omega}^{*} \approx (1-e^2)\tau_{cE}^*/2$, respectively, in the collisional regime, where the prefactor plays an important role, as will be demonstrated later. The consequences of the interplay among these dimensionless numbers will be clarified and discussed in the following sections. Furthermore, we will identify the dimensionless number that -- we believe -- allows us to distinguish between the two scaling regimes. \subsubsection{Simulation parameters} We examine two systems with different grain diameters and composition. The first {\it monodisperse} system consists of only one type of particles, whose diameters are $\sigma_0$. The other {\it polydisperse} system consists of two types of grains, and the diameters of grains are $0.5\sigma_0$, and $\sigma_0$, where the numbers of each type of grains are $0.8N$ and $0.2N$, respectively, with the total number of particles $N$. The reasons to study such a polydisperse system are (i) to avoid crystallization and (ii) to compare our new near-rigid data with previous results from rigid disks \cite{Luding01,Luding02}. In our simulations, the number of particles is $N=2401$ except for the data in Figs.\ \ref{P_eta_poly_e0.9} and \ref{P_eta_poly_log_e0.9}, where we have used $N=20000$. We use the leap-frog algorithm, which is second-order accurate in time, with the time interval $\Delta t=0.2\sqrt{m / k}$. We checked that the simulation converges well by comparison with a shorter time-step $\Delta t=0.02\sqrt{m / k}$. The pressure and the viscosity are respectively given by \begin{eqnarray} P & = & \frac{1}{2V} \left < \sum_{i=1}^N \sum_{j>i} r_{ij} \left [ f_{{\rm el}}(r_{ij}) + f_{{\rm vis}}(r_{ij}, v_{ij,{\rm n}}) \right ] + \sum_{i=1}^N \frac{|\bv{p}_i|^2}{m} \right >, \label{P:ex}\\ \eta & = & -\frac{1}{\dot \gamma V}\left < \sum_{i=1}^N \sum_{j>i} \frac{r_{ij,x} r_{ij,y}}{r_{ij}} \left [ f_{{\rm el}}(r_{ij}) + f_{{\rm vis}}(r_{ij}, v_{ij,{\rm n}}) \right ] + \sum_{i=1}^N \frac{p_{i,x}p_{i,y}}{m} \right > \label{S:calc}, \end{eqnarray} with the volume of the system $V$, the relative distance vector ${\bv{r}}_{ij}=(r_{ij,x},\,r_{ij,y})$, with $r_{ij}=|{\bv{r}}_{ij}|$, and the peculiar momentum $\bv{p}_i = (p_{i,x},\,p_{i,y}) \equiv m(\bv{v}_i - \dot\gamma y_i \bv{e}_x)$. \subsection{Mono-disperse system} \label{Mono:Subsec} In Figs.\ \ref{P_hard}(a) and (b), we plot $P^*$ as a function of the area fraction $\phi$ in the {\it monodisperse} system with $e=0.999$ for $0 < \phi < 0.6$ and $0.5 < \phi < 0.9$, respectively. Most of all data of $P^*$ seem to converge in the rigid-disk limit ($\tau_c^* \to 0$). Moreover, the data for $P^*$ with $\phi<0.6$ are consistent with $P^*_Q$, see Fig.\ \ref{P_hard}(a), while $P^*$ for $\phi>0.7$ in Fig.\ \ref{P_hard}(b) deviates from $P_Q^*$ in the soft case of $\tau_c^*=1.11 \times 10^{-3}$, and also in the rigid-disk limit. Only the simulations with $\tau_c^*=1.11\times 10^{-4}$ are close to $P_Q^*$ -- seemingly by accident. At high densities, for very soft particles, the stress is considerably smaller than predicted by $P^*_Q$, while for near-rigid particles, we observe a higher stress. \begin{figure} \begin{center} \includegraphics[height=14em]{P_hard.eps} \caption{ The reduced pressure $P^*$ as a function of the area fraction $\phi$ in the {\it monodisperse} system with $e=0.999$, for different $\tau_c^*$, as given in the inset, and for $\phi<0.6$ (a) and $\phi > 0.5$ (b). } \label{P_hard} \end{center} \end{figure} \begin{figure} \begin{center} \includegraphics[height=14em]{P_elastic.eps} \caption{ The reduced pressure $P^*$ as a function of the area fraction $\phi$ in the {\it monodisperse} system with $\tau_c^* = 1.11 \times 10^{-5}$ and different $e$, as given in the inset. } \label{P_elastic} \end{center} \end{figure} In order to check the possibility that the restitution coefficient is the reason for the deviation between the numerical data and $P_Q^*$ in Fig.\ \ref{P_hard}(b), we plot $P^*$ for different $e$, for $\tau_c^*=1.11\times 10^{-5}$ in Fig.\ \ref{P_elastic}. At high densities, for inelastically interacting particles, $e=0.99$, the stress is considerably smaller than predicted by $P^*_Q$, while for more elastic particles, we observe a higher stress. Only the almost elastic case $e=0.9999$ is close to the prediction. The low pressure for $e=0.99$ is due to the existence of a shear-band -- see below. For all other situations, no shear-band is observed, however, different patterns of defect lines in the crystal are evidenced for $e=0.9990$ and $e=0.9995$, while an almost perfect crystal is observed for $e=0.9999$, where slip-lines appear. It should be noted that the positions of the slip-lines (shear-bands of width $W=d$) don't move in the steady state of one sample, but vary among different samples. \begin{figure} \begin{center} \includegraphics[height=14em]{grad_snap_mono_e0.99_nu0.84.eps} \caption{ (a) The scaled velocity $u'=u/(\sigma_0 / \sqrt{m/k})$ in $x$ direction as a function of $y'=y/\sigma_0$, for $\phi=0.84$, $\tau_c^* = 1.11 \times 10^{-5}$, and $e=0.99$. (b) Snapshot of the {\it monodisperse} system from (a). } \label{grad_snap_mono_e0.99} \end{center} \end{figure} \begin{figure} \begin{center} \includegraphics[height=14em]{grad_snap_mono_e0.999_nu0.84.eps} \caption{ (a) The scaled velocity $u'$ (like in Fig.\ \ref{grad_snap_mono_e0.99}) for $\phi=0.84$, $\tau_c^* = 1.11 \times 10^{-5}$, and $e=0.999$. (b) Snapshot of the {\it monodisperse} system from (a). } \label{grad_snap_mono_e0.999} \end{center} \end{figure} \begin{figure} \begin{center} \includegraphics[height=14em]{grad_snap_mono_e0.9999_nu0.84.eps} \caption{ (a) The scaled velocity $u'$ (like in Fig.\ \ref{grad_snap_mono_e0.99}) for $\phi=0.84$, $\tau_c^* = 1.11 \times 10^{-5}$, and $e=0.9999$. (b) Snapshot of the {\it monodisperse} system from (a). } \label{grad_snap_mono_e0.9999} \end{center} \end{figure} We have confirmed the existence of shear-bands for $\phi > 0.7$ with $e=0.99$ in Fig.\ \ref{grad_snap_mono_e0.99}. We plot the velocity $u(y)$ in $x$ direction as a function of $y$ for $\phi=0.76$, $\tau_c^* = 1.11 \times 10^{-5}$, and $e=0.99$ in Fig.\ \ref{grad_snap_mono_e0.99} (a), where the velocity gradient exists only in the regions $ y/\sigma_0 < -20$ or $y/\sigma_0 > 20$. The apparent inhomogeneity is observed in the snapshot of the system, see Fig.\ \ref{grad_snap_mono_e0.99}(b). On the other hand, such a shear-band could not be observed for the case of $e=0.999$. Note that the shear-band formation in our system is different from that for the dilute case \cite{Tan} in which dense strips align at 45 degrees relative to the streamwise direction. Fig.\ \ref{grad_snap_mono_e0.999} shows that the system is in an uniformly sheared state with some density fluctuations, see Fig.\ \ref{grad_snap_mono_e0.999}(b). Actually, here deformations take place irregularly and localized -- together with defects and slip planes -- so that the velocity profile looks smooth and linear only after long-time (or ensemble) averaging. For the case of $e=0.9999$, almost perfect crystallization is observed, but slip-lines exist, see Figs.\ \ref{grad_snap_mono_e0.9999}(a) and (b). This is in conflict with the observations of Ref.\ \citen{Luding09}, where shear-bands were observed at densities around $\phi \approx 0.70$, $\phi \approx 0.73$, and $\phi \approx 0.78$, for $e \ge 0.99$, $e=0.95$, and $e=0.90$, respectively. In this paper, for the case of $e=0.999$, no shear band is observed, however, in the simulation of the sheared inelastically interacting rigid-disks with $e=0.998$ in Ref.\ \citen{Luding09}, a shear band was reported. We identify two differences between the systems in this paper and Ref.\ \citen{Luding09}. The first difference is the softness of the disks that, however, should not affect the results as long as we are close to the rigid-disk limit. The second difference is the protocol to obtain a sheared steady state with density $\phi$. In this paper, first an equilibrium state with density $\phi$ is prepared and then shear flow and dissipation between the particles is switched on to obtain the sheared steady state. In contrast, in Ref.\ \citen{Luding09}, the system of sheared inelastically interacting disks was studied by slowly but continuously increasing the density $\phi$. The dimensionless viscosity $\eta^*$ for {\it monodisperse} systems with $e=0.999$, and different $\tau^*_c$ is shown in Fig.\ \ref{eta_hard}. We note that both $P^*$ and $\eta^*$ converge for more rigid disks $\tau_c^* \to 0$, but not to the empirical expression $\eta_L^*$ from Eq.\ \eqref{eta:eq}. It can be used in a wide range of $\phi$, as one can see in Fig.\ \ref{eta_hard}(b), but -- even though behaving qualitatively similar -- the numerical data clearly deviate from $\eta_L^*$: For $\phi > 0.7$, in the rigid-disk case, $\eta_L^*$ diverges at $\phi_\eta=0.71$, whereas $\eta^*$ in the near-rigid case exponentially grows like the Vogel-Fulcher law, which remains finite above $\phi_\eta$. \begin{figure} \begin{center} \includegraphics[height=14em]{eta_hard.eps} \caption{ (a) The dimensionless viscosity $\eta^*$ as a function of the area fraction $\phi$ in the {\it monodisperse} system for $e=0.999$, and different $\tau^*_c$ as given in the inset. (b) $\eta^*/\eta_E^*$ as a function of the area fraction from the same simulations as in (a). } \label{eta_hard} \end{center} \end{figure} The difference between the numerical data for $\eta^*$ and $\eta^*_L$ results from both elasticity and dissipation, as shown in Fig.\ \ref{eta_elastic}, where the dependence of $\eta^*$ on $\phi$ for $\tau_c^* = 1.11 \times 10^{-5}$ and different coefficients of restitution $e$ are plotted. The viscosity $\eta^*$, like the pressure $P^*$, approach $\eta^*_L$ and $P^*_Q$ in the elastic limit $e\to1$, i.e., they converge to the results of the elastic rigid-disk system. It should be noted that Figs.\ \ref{eta_hard}(a) and \ref{eta_elastic}(a) suggest that the singularity around $\phi=\phi_\eta$ is an upper limit, only realized in the rigid disk limit and for $e \rightarrow 1$. As will be discussed below, for given $\tau_c^*$ and $e$, the simulations deviate more and more from the rigid disk case with increasing density. The smaller $\tau_c^*$, i.e., the stiffer the disks, the better is the upper limit approached -- but for finite dissipation and for near-rigid disks, there is always a finite density where the elasticity (softness) becomes relevant and leads to deviations from the upper limit. Above that density, it seems that the divergence of the viscosity takes place at the same point as the pressure, and another inverse power law can be a fitting function for $\phi<\phi_\eta$. \begin{figure} \begin{center} \includegraphics[height=14em]{eta_elastic.eps} \caption{ The dimensionless viscosity $\eta^*$ as a function of the area fraction $\phi$ in the {\it monodisperse} system for $\tau_c^* = 1.11 \times 10^{-5}$ and $e=0.99$, $0.999$, $0.9999$. (b) $\eta^*/\eta_E^*$ as a function of the area fraction from (a). } \label{eta_elastic} \end{center} \end{figure} In rigid-disk systems, the coordination number $Z$ should be identical to zero because the contacts between the particles are instantaneous. Hence, in the rigid-disk limit of soft-disks, it is expected that the coordination number $Z$ vanishes, which is confirmed by Fig.\ \ref{fig:Z}(a). Here, it should be noted that the coordination number $Z$ is almost identical to the dimensionless number $\tau_{cE}^*$ \cite{Luding98}. Indeed the relationship $Z\approx \tau_{cE}^*$ can be verified in Fig.\ \ref{fig:Z}(b), where we plot the ratio $\tau_{cE}^* / Z$ as function of the area fraction $\phi$ for {\it monodisperse} systems with $e=0.999$ and several $\tau_c^*$. Here, we have measured the coordination number as \begin{equation} Z = \sum_i \sum _{j\neq i} \langle \Theta(\sigma_{ij} - r_{ij}) \rangle/N ~. \label{eq:defZ} \end{equation} If we use the mean-field picture, we can understand the relation $Z\approx \tau_{cE}^*$ as shown in Appendix \ref{Z:app}. \begin{figure} \begin{center} \includegraphics[height=14em]{Z.eps} \caption{ (a) The coordination number $Z$ plotted as function of the area fraction $\phi$ for {\it monodisperse} systems with $e=0.999$ for several $\tau_c^*$ values. (b) $\tau_{cE}^* / Z$ plotted as function of the area fraction $\phi$ from the same simulations as in (a). } \label{fig:Z} \end{center} \end{figure} We also show the scaled temperature $T^*$ for the soft-sphere {\it monodisperse} system in Fig.\ \ref{T}. As expected, $T^*$ approaches the empirical expression $T_K^*$ in Eq.\ \eqref{TK}. This result also supports our conjecture that the rigid-disk limit of the soft-disk assemblies coincides with the rigid-disk system when the coefficient of restitution $e$ is sufficiently close to unity. \begin{figure} \begin{center} \includegraphics[height=14em]{T.eps} \caption{ (a) The scaled temperature $T^*$ as a function of the area fraction $\phi$ for the {\it monodisperse} system at $\tau_c^* = 1.11 \times 10^{-5}$ and different $e$. (b) $T^*/T^*_E$ as a function of the area fraction $\phi$ from the same simulations as in (a). \label{T} } \end{center} \end{figure} \subsection{Poly-disperse systems} \label{Poly:Subsec} In order to understand the {\it polydisperse} situation, we also study systems with different $\tau_c^*$ and different $e$ values -- as in the previous subsection. The reduced pressure $P^*$ and the dimensionless viscosity $\eta^*$ are almost independent of $\tau_c^*$ and $e$ for moderate densities ($\phi<0.8$), as shown in Fig.\ \ref{P_eta_poly_wide}, where $P^*$ and $\eta^*$ are plotted as functions of the area fraction $\phi$. For low densities, the simulation results of $P^*$ agree with the scaling given by $P_d^*$, while the asymptotic scaling behavior of $\eta^*$ is described by $\eta_d^*$ only above $\phi \simeq 0.8$. Here, we have used $\phi_{\rm max} = 0.841$ for $P^*$ and $\eta^*$ in Eqs.\ \eqref{Pd:eq} and \eqref{etad:eq}. \begin{figure} \begin{center} \includegraphics[height=14em]{P_eta_poly_wide.eps} \caption{ (a) The dimensionless pressure $P^*$ as a function of the area fraction $\phi$ for {\it polydisperse} systems with several different $\tau^*_c$ and $e$, where we have used $\phi_{\rm max} = 0.841$ for $P_d^*$ and $\eta^*_d$. The prefactor for $P_d^* \propto (\phi_{\rm max} - \phi)^{-1}$ is chosen as $2\phi_{\rm max}$ \cite{Luding01,Luding02,Luding09}. (b) The dimensionless viscosity $\eta^*$ as a function of the area fraction $\phi$ from the same simulations as in (a). Here, we have used the prefactor 7.0 for $\eta_d^* \propto (\phi_{\rm max} - \phi)^{-1}$. } \label{P_eta_poly_wide} \end{center} \end{figure} However, when looking more closely, there are distinct differences between $P^*$ and $P_d^*$, and between $\eta^*$ and $\eta^*_d$ for $\phi > 0.83$. In Fig.\ \ref{P_eta_poly_e0.9}, $P^*$ and $\eta^*$ are plotted from {\it polydisperse} systems with rather strong dissipation, $e=0.9$, where we have used particular values for $\phi_{\rm max} = 0.841$ and $\phi_J = 0.8525$ in order to visualize their different behavior. Although $P^*$ is still finite for $\phi > \phi_{\rm max}$ in the hard disk limit, even for the smallest $\tau^*_c$ values, both $P_d^*$ and $\eta_d^*$ diverge at $\phi_{\rm max}$ as $(\phi_{\rm max} - \phi)^{-1}$. On the other hand, in the same high density range, $P^*$ and $\eta^*$ are consistent with $P_J^*$ \eqref{scaling:P} and $\eta_J^*$ \eqref{scaling:eta}~\cite{Otsuki:PTP,Otsuki:PRE} in the rigid-disk limit ($\tau_c^* = 1.11 \times 10^{-6}$), as will be shown below. \begin{figure} \begin{center} \includegraphics[height=14em]{P_eta_poly_e0.9.eps} \caption{ (a) The dimensionless pressure $P^*$ as a function of the area fraction $\phi$ for the {\it polydisperse} system for $e=0.9$ and several $\tau^*_c$. (b) The dimensionless viscosity $\eta^*$ as a function of the area fraction $\phi$ in the {\it polydisperse} system from the same simulations as those in (a). Here, we have used $\phi_{J}=0.8525$ for $P^*_J$ and $\eta^*_J$, and $\phi_{\rm max} = 0.841$ for $P^*_d$ and $\eta^*_d$. The prefactors for $P_d^* \propto (\phi_{\rm max} - \phi)^{-1}$ and $\eta_d^* \propto (\phi_{\rm max} - \phi)^{-1}$ are $2\phi_{\rm max}$ and $7.0$, respectively. For $P_J^* \propto (\phi_{J} - \phi)^{-2}$ and $\eta_J^* \propto (\phi_{J} - \phi)^{-3}$, the prefactors are chosen as $0.07$ and $0.002$, respectively. } \label{P_eta_poly_e0.9} \end{center} \end{figure} In order to verify whether the critical behavior of $P^*$ and $\eta^*$ can be described by $P_J^*$ and $\eta_J^*$, we plot $P^*$ and $\eta^*$ as functions of $\phi_J - \phi$ in Fig.\ \ref{P_eta_poly_log_e0.9}. Here, we plot only the data for $\phi < \phi_J$ because we discuss the scaling behavior of $P^*$ and $\eta^*$ in the unjammed regime in this paper. $P^*$ and $\eta^*$ in the rigid-disk limit approach $P_J^*$ and $\eta_J^*$, which satisfy $(\phi_J - \phi)^{-2}$ and $(\phi_J - \phi)^{-3}$, respectively. It should be noted that the plateaus in Fig.\ \ref{P_eta_poly_log_e0.9}, close to the jamming transition point, for $\phi \simeq \phi_J$, can also be predicted from the scaling theory, by rewriting Eqs.\ \eqref{scaling}--\eqref{exponents}. More specifically, the arguments are taken to the power $-1/\alpha$: \begin{equation} T = \dot \gamma^{x_{\Phi}/\alpha} {\cal T}'_{\pm} \left(\frac{|\Phi|}{\dot\gamma^{1/\alpha}}\right), \ S = \dot \gamma^{y_{\Phi}/\alpha}{\cal S}'_{\pm} \left(\frac{|\Phi|}{\dot\gamma^{1/\alpha}}\right), \ P = \dot \gamma^{y_{\Phi}'/\alpha}{\cal P}'_{\pm} \left(\frac{|\Phi|}{\dot\gamma^{1/\alpha}}\right), \end{equation} where we have introduced ${\cal T}'_{\pm}(x) = x^{-x_\Phi}{\cal T}_{\pm}(x^{-\alpha})$, ${\cal S}'_{\pm}(x) = x^{-y_\Phi}{\cal S}_{\pm}(x^{-\alpha})$, and ${\cal P}'_{\pm}(x) = x^{-y'_\Phi}{\cal P}_{\pm}(x^{-\alpha})$. The scaling functions satisfy $\lim_{x\to 0}{\cal T}'_{\pm}(x)=\lim_{x\to 0}{\cal S}'_{\pm}(x)=\lim_{x\to 0}{\cal P}'_{\pm}(x) = const.$ Substituting these relations into Eqs.\ \eqref{P:def} \eqref{eta:def}, with Eqs.\ \eqref{exponents}, $\eta=S/\dot\gamma$, $\Delta=1$, and the definition of $\tau_c^*$ given by Eq.\ \eqref{tau:def}, the scaling relations of $P^*$ and $\eta^*$ are obtained as \begin{equation} P^* = \tau_c^{*-4/5} {\cal P}^*_{\pm} \left(\frac{|\Phi|}{\tau_c^{*2/5}}\right), \ \eta^* = \tau_c^{*-6/5} {\cal H}^*_{\pm} \left(\frac{|\Phi|}{\tau_c^{*2/5}}\right). \end{equation} Here, the scaling functions satisfy $\lim_{x\to 0}{\cal P}^*_{\pm}(x)=\lim_{x\to 0}{\cal H}^*_{\pm}(x) = const.$ Therefore, the plateau for $P^*$ and $\eta^*$ in Fig.\ \ref{P_eta_poly_log_e0.9} should be proportional to $(1/\tau_c^*)^{4/5}$ and $(1/\tau_c^*)^{6/5}$, respectively, which is confirmed by Fig.\ \ref{scale_P_eta}, where we plot $P^* \tau_c^{*4/5}$ and $\eta^* \tau_c^{*6/5}$ as a function of $(\phi_J-\phi)/\tau_c^{*2/5}$. \begin{figure} \begin{center} \includegraphics[height=14em]{P_eta_poly_log_e0.9.eps} \caption{ (a) The reduced pressure $P^*$ plotted as a function of $\phi_J - \phi$ for {\it polydisperse} systems with $e=0.9$ and several $\tau^*_c$ based on the simulations used for Fig.\ \ref{P_eta_poly_e0.9}. (b) The dimensionless viscosity $\eta^*$ from the same simulations as those in (a). Here, we have used $\phi_{J}=0.8525$ for $P^*_J$ and $\eta^*_J$, and $\phi_{\rm max} = 0.841$ for $P^*_d$ and $\eta^*_d$. } \label{P_eta_poly_log_e0.9} \end{center} \end{figure} \begin{figure} \begin{center} \includegraphics[height=14em]{scale_P_eta.eps} \caption{ (a) Plots of $P^* \tau_c^{*4/5}$ versus $(\phi_J-\phi)/\tau_c^{*2/5}$ for {\it polydisperse} systems with $e=0.9$ and several $\tau^*_c$. (b) Plots of $\eta^* \tau_c^{*6/5}$ versus $(\phi_J-\phi)/\tau_c^{*2/5}$ for {\it polydisperse} systems with $e=0.9$ and several $\tau^*_c$. } \label{scale_P_eta} \end{center} \end{figure} Whether the simulation pressure is described by $P_d^*$ or $P_J^*$, and whether the viscosity is given by $\eta_d^*$ or $\eta_J^*$, strongly depends on the coefficient of restitution $e$. In Figs.\ \ref{P_eta_poly_e0.1}--\ref{P_eta_poly_e0.998}, we plot $P^*$ and $\eta^*$ as functions of $\phi$ for various $e$, involving the very high dissipation case $e=0.1$, an intermediate case $e=0.99$, and a low dissipation case $e=0.998$. Using fitting values $\phi_{\rm max}=0.841$, $0.848$, and $0.851$, based on a fit starting from very low densities, corresponding to various $e=0.1$, $0.99$ and $0.998$, respectively, we can approximate the data of $P^*$ best by $P_d^* = 2 \phi_{\rm max}/(\phi_{\rm max} - \phi)$. On the other hand, we assume that $\phi_J$ is independent of $e$, and fix $\phi_J=0.8525$ for all $e$, as confirmed this by our numerical simulations. Even in the case of strong inelasticity ($e=0.1$), as shown in Fig.\ \ref{P_eta_poly_e0.1}, $P_J^*$ and $\eta_J^*$ characterize the behavior of $P^*$ and $\eta^*$ near the jamming transition point, while $P_d^*$ and $\eta_d^*$ deviate for $\phi > 0.83$. The range where $P_J^*$ and $\eta_J^*$ characterize the pressure and the viscosity becomes narrower as $e \to 1$, while the range of validity of $P_d^*$ becomes wider, as shown in Figs.\ \ref{P_eta_poly_e0.99} and \ref{P_eta_poly_e0.998}. For $e=0.998$ (Fig.\ \ref{P_eta_poly_e0.998}), the difference between $P_d^*$ and $P_J^*$ appears only in a small region of $\phi$ which is shown in Fig.\ \ref{P_eta_poly_log_e0.998}. Since the scaling behaviors of $P^*$ and $\eta^*$ agree with $P_J^*$ and $\eta_J^*$ near $\phi_J$, we conclude that the critical behavior for inelastic near-rigid systems is well described by $P_J^*$ and $\eta_J^*$, as proposed in Refs.~\citen{Otsuki:PTP,Otsuki:PRE}. The scaling plot in Fig.\ \ref{scale_P_eta} supports the validity of the critical behaviors concerning both the plateaus and the lower densities. However, such predictions cannot be used for almost elastic and perfectly elastic systems, neither mono- or polydisperse, whose critical behavior is described by $P_d^*$ and $\eta_d^*$ instead. \begin{figure} \begin{center} \includegraphics[height=14em]{P_eta_poly_e0.1.eps} \caption{ (a) The reduced pressure $P^*$ as a function of the area fraction $\phi$ for the {\it polydisperse} system with $e=0.1$ and several $\tau^*_c$. (b) The dimensionless viscosity $\eta^*$ from the same data as those in (a). We have used $b=0.07$ and $\phi_J=0.8525$ for $P_J$ and $\eta_J$, and $\phi_{\rm max} = 0.841$ for $P_d$ and $\eta_d$. We used $\phi_{J}=0.8525$ for $P^*_J$ and $\eta^*_J$, and $\phi_{\rm max} = 0.841$ for $P^*_d$ and $\eta^*_d$ The prefactors for $P_d^* \propto (\phi_{\rm max} - \phi)^{-1}$ and $\eta_d^* \propto (\phi_{\rm max} - \phi)^{-1}$ are $2\phi_{\rm max}$ and $7.0$, respectively. The prefactors for $P_J^* \propto (\phi_{J} - \phi)^{-2}$ and $\eta_J^* \propto (\phi_{J} - \phi)^{-3}$ are given by $0.07$ and $0.002$, respectively. } \label{P_eta_poly_e0.1} \end{center} \end{figure} \begin{figure} \begin{center} \includegraphics[height=14em]{P_eta_poly_e0.99.eps} \caption{ (a) The reduced pressure $P^*$ as a function of the area fraction $\phi$ for the {\it polydisperse} system with $e=0.99$ and several $\tau^*_c$. (b) The dimensionless viscosity $\eta^*$ obtained from the same data as those in (a). We have used $\phi_J=0.8525$ for $P_J$ and $\eta_J$, and $\phi_{\rm max} = 0.848$ for $P_d$ and $\eta_d$. The prefactors for $P_d^* \propto (\phi_{\rm max} - \phi)^{-1}$ and $\eta_d^* \propto (\phi_{\rm max} - \phi)^{-1}$ are $2\phi_{\rm max}$ and $10.0$, respectively. The prefactors for $P_J^* \propto (\phi_{J} - \phi)^{-2}$ and $\eta_J^* \propto (\phi_{J} - \phi)^{-3}$, are $0.035$ and $0.0015$, respectively. } \label{P_eta_poly_e0.99} \end{center} \end{figure} \begin{figure} \begin{center} \includegraphics[height=14em]{P_eta_poly_e0.998.eps} \caption{ (a) The reduced pressure $P^*$ as a function of the area fraction $\phi$ for the {\it polydisperse} system for $e=0.998$ and several $\tau^*_c$. (b) The dimensionless viscosity $\eta^*$ obtained from the same data as those in (a). We have used $\phi_J=0.8525$ for $P^*_J$ and $\eta_J$, and $\phi_{\rm max} = 0.851$ for $P^*_d$ and $\eta_d$. The prefactors for $P_d^* \propto (\phi_{\rm max} - \phi)^{-1}$ and $\eta_d^* \propto (\phi_{\rm max} - \phi)^{-1}$ are $2\phi_{\rm max}$ and $25.0$, respectively. The prefactors for $P_J^* \propto (\phi_{J} - \phi)^{-2}$ and $\eta_J^* \propto (\phi_{J} - \phi)^{-3}$, are $0.01$ and $0.001$, respectively. } \label{P_eta_poly_e0.998} \end{center} \end{figure} \begin{figure} \begin{center} \includegraphics[height=14em]{P_eta_poly_log_e0.998.eps} \caption{ (a) The reduced pressure $P^*$ plotted as a function of $\phi_J - \phi$ for {\it polydisperse} systems with $e=0.998$ and several $\tau^*_c$ based on the simulations used for Fig.\ \ref{P_eta_poly_e0.998}. (b) The dimensionless viscosity $\eta^*$ obtained from the same simulations as those in (a). We have used $\phi_J=0.8525$ for $P^*_J$ and $\eta_J$, and $\phi_{\rm max} = 0.851$ for $P^*_d$ and $\eta_d$. The prefactors for $P_d^* \propto (\phi_{\rm max} - \phi)^{-1}$ and $\eta_d^* \propto (\phi_{\rm max} - \phi)^{-1}$ are $2\phi_{\rm max}$ and $25.0$, respectively. The prefactors for $P_J^* \propto (\phi_{J} - \phi)^{-2}$ and $\eta_J^* \propto (\phi_{J} - \phi)^{-3}$, are $0.01$ and $0.001$, respectively. } \label{P_eta_poly_log_e0.998} \end{center} \end{figure} \subsection{Dimensionless numbers and a criterion for the two scaling regimes} \label{Time:Subsec} In Sec.\ \ref{Poly:Subsec}, we reported a crossover from the region satisfying Eqs.\ (\ref{Pd:eq}) and (\ref{etad:eq}) to the region satisfying Eqs.\ (\ref{scaling:P}) and (\ref{scaling:eta}). Figure \ref{phase} presents a schematic phase diagram in the plane of the restitution coefficient $e$ and the area fraction $\phi$, where -1 denotes the region satisfying the scaling relations given by Eqs.\ (\ref{Pd:eq}) and (\ref{etad:eq}), and OH denotes the region satisfying the scalings given by Eqs.\ (\ref{scaling:P}) and (\ref{scaling:eta}). For each $e$, the high density region satisfies Eqs.\ (\ref{scaling:P}) and (\ref{scaling:eta}), while the low density region satisfies the scalings given by Eqs.\ (\ref{Pd:eq}) and (\ref{etad:eq}). As the restitution coefficient approaches unity, the region of OH becomes ``narrower'', and disappears in the elastic limit. \begin{figure} \begin{center} \includegraphics[height=14em]{phase.eps} \caption{ A schematic phase diagram of the region (-1) satisfying Eqs.\ (\ref{Pd:eq}), (\ref{etad:eq}) and the region (OH) satisfying Eqs.\ (\ref{scaling:P}), (\ref{scaling:eta}). } \label{phase} \end{center} \end{figure} Now, let us discuss which of the dimensionless numbers $\tau_E^*$, $\tau_{cE}^*$, $\tau_\omega^*$ or $\tau_{c\omega}^{*}$ can be used as the criterion to distinguish between the two scaling regimes. It should be noted that the dimensionless number for the criterion must be a monotonic function of $\phi$, because the scaling relations Eqs.\ (\ref{Pd:eq}) and (\ref{etad:eq}) appear in the higher density region and the scaling relations Eqs.\ (\ref{scaling:P}) and (\ref{scaling:eta}) appear in the lower density region regardless to other parameters. First, let us consider $\tau_E^*$. We expect that $\tau_E^* < A$ or $\tau_E^* > A$ is the criterion for the scaling regime given by (\ref{scaling:P}) and (\ref{scaling:eta}), where $A$ is a constant. However, since $\tau_E^*$ is not a monotonic function of the area fraction $\phi$ and the restitution coefficient $e$, as shown in Fig.\ \ref{tauE_poly_elastic}(a), we conclude that neither $\tau_E^* < A$ or $\tau_E^* > A$ is appropriate for the criterion. Similar to the case of $\tau_{E}^*$, $\tau_{cE}^*$ in not a monotonic function of $\phi$ and $e$, as shown in Fig.\ \ref{tauE_poly_elastic}(b). Therefore, we conclude that $\tau_{cE}^*$ is not an appropriate dimensionless time for the criterion. \begin{figure} \begin{center} \includegraphics[height=14em]{tauE_poly_elastic.eps} \caption{ (a) $\phi$ dependence on $\tau_{E}^*$ and (b) $\phi$ dependence on $\tau_{cE}^*$, for various $e$, from simulations with $\tau_c^*=1.1 \times 10^{-5}$. } \label{tauE_poly_elastic} \end{center} \end{figure} Finally, let us consider $\tau_{\omega}^*$ and $\tau_{c\omega}^*$, which are respectively related with $\tau_E^*$ and $\tau_{cE}^*$ as $\tau_\omega^* \approx 2\tau_{E}^* / (1-e^2)$ and $\tau_{c\omega}^{*}\approx (1-e^2)\tau_{cE}^*/2$ in the collisional regime, but their dependency on $\phi$ and $e$ differs from those of $\tau_E^*$ and $\tau_{cE}^*$, as shown in Figs.\ \ref{tauw_poly_elastic}(a) and \ref{tauw_poly_elastic}(b). Both $\tau_{\omega}^*$ and $\tau_{c\omega}^*$ are monotonic functions of $\phi$ and $e$. Since Eqs.\ (\ref{scaling:P}) and (\ref{scaling:eta}) are satisfied in the high density region and $\tau_\omega^*$ and $\tau_{c\omega}^*$ are respectively decreasing and increasing functions of the density $\phi$, $\tau_\omega^*<A$ and $\tau_{c\omega}^*>A$ are the possible conditions for the scaling given by Eqs.\ (\ref{scaling:P}) and (\ref{scaling:eta}). These conditions are also consistent with the dependencies of $\tau_\omega^*$ and $\tau_{c\omega}^*$ on $e$. Indeed, $\tau_\omega^*$ increases as the restitution constant increases, and $\tau_{c\omega}^{*}$ is a decreasing function of $e$. This means that the regions satisfying $\tau_\omega^*<A$ and $\tau_{c\omega}^*>A$ are narrower as the restitution constant increases, which is consistent with the numerical observation. Therefore, $\tau_\omega^*<A$ and $\tau_{c\omega}^*>A$ are the only two possible candidates to characterize the system with respect to their scaling behavior. It should be noted that $\tau_{c\omega}^*$ tends to zero in the hard disk limit $\tau_c^* \to 0$. In this sense, to use $\tau_{c\omega}^*$ might involve a conceptual difficulty, even though $\tau_{c\omega}^*$ is finite in the jamming region. \begin{figure} \begin{center} \includegraphics[height=14em]{tauw_poly_elastic.eps} \caption{ (a) $\phi$ dependence on $\tau_{\omega}^*$ and (b) $\phi$ dependence on $\tau_{c\omega}^{*}$ for various $e$, from simulations with $\tau_c^*=1.1 \times 10^{-5}$. } \label{tauw_poly_elastic} \end{center} \end{figure} \section{Conclusion and Discussion} \label{Conclusion:Sec} In conclusion, we have investigated the dimensionless pressure $P^*$ and the dimensionless viscosity $\eta^*$ of two-dimensional soft disk systems and have payed special attention to the rigid-disk limit of inelastically interacting systems, while near-rigid disks still have some elasticity (``softness''). For {\it monodisperse} systems, as the system approaches the elastic limit, $e \to 1$, both $P^*$ and $\eta^*$ for $\phi<\phi_\eta=0.71$ approach the results of elastic rigid-disk systems, where the viscosity increases rapidly around $\phi=\phi_\eta$ due to ordering (crystallization) effects, while the pressure for $\phi>\phi_\eta$ is still finite \cite{Garcia}. This result is consistent with Ref.\ \citen{Mitarai}, where Mitarai and Nakanishi suggested that the behavior of soft-disks in dilute collisional flow converges to that of rigid-disks in the rigid-disk limit. For {\it polydisperse} systems, both $P^*$ and $\eta^*$ behave as $(\phi_J - \phi)^{-2}$ and $(\phi_J - \phi)^{-3}$ near the jamming transition point, $\phi_J > \phi_\eta$, as predicted in Refs.~\citen{Otsuki:PTP,Otsuki:PRE}. However, as the restitution coefficient $e$ approaches unity, the scaling regime becomes narrower, and the exponents for the divergence of $P^*$ and $\eta^*$ approach values close to $-1$ in the almost elastic case. From these results, we conclude that the predictions for the inelastic soft-disk systems in Refs.~\citen{Otsuki:PTP,Otsuki:PRE} are applicable to the inelastic near-rigid disk systems below the jamming transition point, but the prediction cannot be used for almost elastic rigid-disk systems. It seems that $\tau_{c\omega}^{*}$ and $\tau_\omega^*$ are the only two possible candidates to characterize the criterion of this crossover. In other words, the energy dissipation rate and the shear rate set the two competing time-scales that define the dimensionless number $\tau_\omega^*$. For $\tau_\omega^* \ll 0.01$ the near-rigid, dissipative scaling regime occurs, while for $\tau_\omega^* \gg 0.01$ the rigid, elastic scaling regime is realized. In three-dimensional sheared inelastic soft-sphere systems \cite{Otsuki:PTP,Otsuki:PRE}, even in {\it monodisperse} cases, there is no indication of the strong ordering transition, and the scaling given in Eqs.\ \eqref{scaling:P} and \eqref{scaling:eta} seems to be valid. However, a direct comparison of near-rigid sphere with rigid sphere simulations in the spirit of the present study is unavailable to our knowledge. We restricted our interest to frictionless particles. When the particles have friction, the scaling relations for the divergence of the viscosity and the pressure may be different, as will be discussed elsewhere. Furthermore, the very soft or high shear rate regime also needs further attention in both 2D and 3D. \section*{Acknowledgements} This work was supported by the Grant-in-Aid for scientific research from the Ministry of Education, Culture, Sports, Science and Technology (MEXT) of Japan (Nos.~21015016, 21540384, and 21540388), by the Global COE Program ``The Next Generation of Physics, Spun from Universality and Emergence'' from MEXT of Japan, and in part by the Yukawa International Program for Quark-Hadron Sciences at Yukawa Institute for Theoretical Physics, Kyoto University. The numerical calculations were carried out on Altix3700 BX2 at YITP in Kyoto University. SL acknowledges the hospitality at YITP in Kyoto, and support from the Stichting voor Fundamenteel Onderzoek der Materie (FOM), financially supported by the Nederlandse Organisatie voor Wetenschappelijk Onderzoek (NWO).
\section{Introduction} This manuscript addresses a couple of open problems in the Skyrme-Hartree-Fock (SHF) scheme. Before coming to these puzzling details, we want to emphasize the enormous merits of SHF. The SHF energy functional manages to establish a reliable description of nuclear properties all over the chart of isotopes (except perhaps the lightest ones) with an adjustment of only a dozen universal parameters, for reviews see e.g. \cite{Ben03aR,Stone_rew}. The enormous success of SHF implies the temptation to ask for more details and it is mostly here where we encounter the present limitations of SHF. The aim of this contribution is to identify problems in order to solve them later on in a common effort of the nuclear physics community. We will in the following address five points: the interpretation as ``force'', the consistency of ground state correlations, the giant dipole resonance in light nuclei, extrapolation of binding energies to super-heavy elements (SHE), and fission of SHE. The first two points are of formal nature, the last three more phenomenological. \section{Formal inconsistencies} \subsection{The concept of a ``force''} One often thinks in terms of a ``Skyrme-force'', to be more precise a Skyrme interaction, whose most touchy ingredient is the density dependent term \begin{equation} \hat{V}_3(\mathbf{r}_1,\mathbf{r}_2) = \frac{t_3}{6}\delta(\mathbf{r}_1-\mathbf{r}_2) \rho^\alpha(\mathbf{r}_1) \left(1+x_3\hat{\Pi}_{\sigma}\right) \quad,\quad \rho(\mathbf{r}) = \langle\Phi|\hat\rho(\mathbf{r})|\Phi\rangle \quad, \label{eq:3body} \end{equation} where $\hat{\Pi}_{\sigma}$ is the spin-exchange operator. We argue that this is a most dubious object. It is not a ``stand alone'' interaction operator, but depends on a mean-field state $|\Phi\rangle$ from which the density $\rho(\mathbf{r})$ is taken. Non-integer values of $\alpha$ immediately hinder an identification with an $N$-body force. The simplest case is $\alpha=1$ and one is tempted to interpret the interaction then as a three body force. Let us consider this case and also ignore the spin-exchange term by setting $x_3=0$ for simplicity. If the operator (\ref{eq:3body}) was a true interaction operator, one should be able to produce an equivalent expression in terms of Fermion operators as \begin{equation} \hat{V}_3^\mathrm{(FO)} = \sum_{\alpha_1\alpha_2\alpha_3\beta_1\beta_2\beta_3} V_{\alpha_1\alpha_2\alpha_3\beta_1\beta_2\beta_3} \hat{a}^\dagger_{\alpha_1} \hat{a}^\dagger_{\alpha_2} \hat{a}^\dagger_{\alpha_3} \hat{a}^{\mbox{}}_{\beta_3} \hat{a}^{\mbox{}}_{\beta_2} \hat{a}^{\mbox{}}_{\beta_1} \quad. \label{eq:V3-FO} \end{equation} The ground-state expectation value of such a $\hat{V}_3^\mathrm{(FO)}$ reads \begin{equation} \langle\Phi|\hat{V}_3^\mathrm{(FO)}|\Phi\rangle = \sum_{nmk}V_{nmk,\widetilde{nmk}} \quad, \label{eq:V3-expFO} \end{equation} where $\widetilde{nmk}$ stands for anti-symmetrization of all three states $nmk$. However, the expectation value of the effective interaction (\ref{eq:3body}) for $\alpha=1$ reads \begin{eqnarray} \langle\Phi|\hat{V}_3|\Phi\rangle &\propto& \sum_{nmk=1}^N \int d^3r \left[ \varphi^\dagger_{n} \varphi_{n}^{\mbox{}} \varphi^\dagger_{m} \varphi_{m}^{\mbox{}} - \varphi^\dagger_{n} \varphi_{m}^{\mbox{}} \varphi^\dagger_{m} \varphi_{n}^{\mbox{}} \right] \varphi^\dagger_{k} \varphi_{k}^{\mbox{}} \nonumber\\ &\equiv& \sum_{nmk}V_{nmk,\widetilde{nm}\,k} \quad. \label{eq:V3-effexp} \end{eqnarray} Note that the state $k$ is not included in the anti-symmetrization. Thus the whole expression can never be written in the form (\ref{eq:V3-expFO}) and the Skyrme ansatz (\ref{eq:3body}) cannot be interpreted as an interaction. A unique and consistent object is the total energy which turns out to be a functional of the local density $\rho(\mathbf{r})$ and spin density ${\boldsymbol\sigma}(\mathbf{r})$, i.e. for $\alpha=1$ and $x_3=0$ \begin{eqnarray} E_3 &=& \langle\Phi|\hat{V}_3|\Phi\rangle = \frac{t_3}{24} \int d^3r\Big\{ 2\rho^{3} - \rho\left(\rho_n^2+\rho_p^2\right) - \rho\left({\boldsymbol\sigma}^2_p+{\boldsymbol\sigma}^2_n\right) \Big\} \quad. \label{eq:E3-funct} \end{eqnarray} The main use of the interaction (\ref{eq:3body}) is that it serves nicely as a formal generator for that functional. But any other use is dangerous. Let us consider, e.g., the residual interaction in RPA. It is deduced from the energy functional by second functional derivative \cite{Rei-RPA} and reads, e.g., for pure density variations (no spin excitations) \begin{equation} V^\mathrm{(res)}_3 = \frac{\partial^2E_3} {\partial\rho(\mathbf{r_1})\partial\rho(\mathbf{r_2})} = \frac{t_3}{2} \delta(\mathbf{r}_1-\mathbf{r}_2) \rho(\mathbf{r}_1) \quad. \end{equation} This has a strength factor $t_3/2$ which is different from the $t_3/6$ of the initial interaction (\ref{eq:3body}). Thus the interaction (\ref{eq:3body}) is not consistently reproducible by standard many-body techniques. Most energy-density functionals are plagued by the self-interaction error \cite{Drei90aB}. It can be checked simply by considering the case of exactly one particle. Functionals with self interaction then yield still a non-vanishing energy which is, of course, unphysical. The functional (\ref{eq:E3-funct}) yields correctly value zero for the case of one particle and is thus self-interaction free. That nice feature is achieved by derivation from the interaction (\ref{eq:3body}). It ought to be mentioned, however, that a derivation from an interaction is not a necessary condition for constructing self-interaction free energy functionals. We thus have seen from two different aspects that the notion of a Skyrme ``force'' is misleading. The cleanest view of Skyrme-Hartree-Fock is to derive it from an energy-density functional. On the other hand, deriving the functional (\ref{eq:E3-funct}) from the effective interaction (\ref{eq:3body}) avoids the self-interaction error and provides the spin terms which otherwise would be much undetermined. It is a matter of phenomenology to check whether the thus imposed spin terms in the nuclear energy density functional are supported by phenomenological data. \subsection{Fragmentation and collective correlations} The ground state of a nucleus is usually computed with a correction of the center-of-mass energy. The motivation is that the mean-field state violates translational invariance and that one needs to consider an ``intrinsic'' state which is obtained by center-of-mass projection. This projection can be simplified by many-body techniques (second order Gaussian-Overlap-Approximation \cite{Rei-cm}) to \begin{equation} E_\mathrm{cm} = \frac{\langle\Phi|\hat{P}_\mathrm{cm}^2|\Phi\rangle}{2mA} \approx 30\,\mathrm{MeV}\,A^{-1/3} \quad. \label{eq:Ecm} \end{equation} Both forms (the operator expectation value or the simple estimate) are widely used and both include the total nucleon number $A$. Now consider fusion of two nuclei. Initially, we have a c.m. energy (\ref{eq:Ecm}) for each nucleus $A_1$ and $A_2$, but finally only one for the total $A$, i.e. \begin{equation*} E_\mathrm{cm}^\mathrm{(in)} = E_\mathrm{cm}(A_1) + E_\mathrm{cm}(A_2) \quad\stackrel{?}{\longleftrightarrow}\quad E_\mathrm{cm}^\mathrm{(fus)} = E_\mathrm{cm}(A_1+A_2) \quad. \end{equation*} That is inconsistent and is particularly puzzling in between where one does not know which one of the both rules to apply. The problem was already noted in \cite{Ber80,Ska07} and an interpolation formula was proposed as an ad-hoc remedy. We want to analyze the case further. Short closer inspection shows that the six initial c.m. degrees-of-freedom merge into three final c.m. degrees-of-freedom, two rotational degrees-of-freedom (for axial symmetry of the final state), and one quadrupole mode. \begin{figure}[h!] \centerline{ \epsfig{figure=figure1.eps,width=0.60\linewidth} \vspace*{-1cm} } \caption{\label{fig:E_cm} Ambiguity of collective degrees of freedom illustrated for an axial symmetric fusion process of two spherical nuclei. The initial six collective c.m. modes ($\{P^{(i)}_x,P^{(i)}_y,P^{(i)}_z\}_{i=1,2}$) merge into three c.m. modes ($\{P_x,P_y,P_z\}$) of the compound, two rotational modes ($\{J_x,J_y\}$) and a quadrupole vibration mode ($\{P_{20}\}$). } \end{figure} The problem could be resolved by associating (axial) quadrupole and rotational correlations with the compound nucleus. However, {this imposes} a new problem: we should do the same with the initial two nuclei. This, in turn, provides even more initial degrees-of-freedom (12 instead of 6) which have to merge into further collective modes of the compound system. This loop generates more and more correlating modes and it is not clear where to stop. The problem may be bearable as long as one considers only intact nuclei with fixed particle number. It becomes a big hindrance in any reaction which changes particle number. Thus there is an urgent need to develop a counting of collective correlations which is robust under fission, fusion and fragmentation. For the time being, it is the most consistent procedure to assume that all correlations are already built into the energy-density functional and to discard any correlation correction, even the ones for c.m. or rotational motion. That holds particularly for all TDHF calculations of large amplitude collective motion. \section{Trend of the GDR with mass number} Giant resonances are crucial nuclear excitation modes. They can be described consistently with a given energy functional by using time-dependent density functional theory, in the nuclear context called TDHF, and considering the small amplitude limit thereof. The scheme is called Random-Phase-Approximation (RPA), for details see \cite{Rei-RPA}. Information from giant resonances in heavy nuclei has often been used in the calibration of a Skyrme parameterization, see e.g. \cite{Brack-skms}. A particular prominent mode is the Giant Dipole Resonance (GDR) which is commonly believed to be well under control with SHF. However, this holds only for heavy nuclei. \begin{figure} \centerline{ \epsfig{figure=figure2.eps,width=0.95\linewidth} } \caption{\label{fig:GDR_trends_comp_scale-2} Peak energies of the Giant-Dipole-Resonance (GDR) drawn versus inverse radius $R=1.16\,\mathrm{fm}\,A^{1/3}$. Heavy nuclei(e.g. Pb) are found at the left side and light nuclei (e.g. O) to the right. The energies are scaled with $A^{1/6}$. Compared are results from two different Skyrme parameterizations and experimental data. Left: Peak energy from a full RPA calculation. Right: Energy from a sum-rule estimate. } \end{figure} This is demonstrated in the left panel of figure \ref{fig:GDR_trends_comp_scale-2} which compares RPA values for the average energy of the GDR with experimental data. The peak energies are deduced from the dipole strength distributions. We show results for two Skyrme forces (SkM$^*$ \cite{Brack-skms} and SLy6 \cite{sly46}). We have checked a broad variety of other Skyrme forces and {always find} the same trend. The discrepancy is obvious: while the GDR for heavy nuclei can be adjusted very well, it is impossible to have simultaneously a reasonable description in small nuclei. The trend is grossly wrong. The experimental data comply fairly well with a trend $E_\mathrm{GDR}\propto A^{-1/6}$, but RPA predicts a much different trend with a sizeable admixture of $E_\mathrm{GDR}\propto A^{-1/3}$. The right panel of figure \ref{fig:GDR_trends_comp_scale-2} shows results of an estimate using the Thomas-Reiche-Kuhn (TRK) sum rule \cite{Rei-RPA,Brack-GDR}. The peak energy is, of course, overestimated. But the trend $\propto A^{-1/6}$ complies with experiment. The TRK mode is a surface mode (Goldhaber-Teller). The competitor is the volume mode (Steinwedel-Jensen) which produces a trend $\propto A^{-1/3}$ \cite{Brack-GDR}. We thus see that the RPA description underestimates the surface contribution and leaves to much bias on the volume. The conjecture is that the present Skyrme forces are still having an inappropriate isovector surface energy. Substantial improvement in that part is needed. \section{Extrapolation to SHE} \begin{SCfigure}[0.8][h!] { \epsfig{figure=figure3.eps,width=9cm}} \caption{\label{fig:SV-min-def-energies} Deviation from the experimental binding energies for a SHF calculation with subsequent correlation corrections for all available nuclei drawn versus mass number $A$. The nuclei which were included in the fit of SV-min, or SV-def respectively, are indicated by filled boxes, well deformed nuclei ($\beta_2>0.2$) by grey circles, and all others by small triangles. Upper: Results from SV-min. Lower: Results from SV-def, a force with $^{264}$Hs added to the fit data. The experimental binding energies were taken from \cite{Aud03}. } \end{SCfigure} Skyrme parameterizations are usually determined by a phenomenological adjustment to a given pool of fit data (binding energies, radii, etc., for a chosen set of nuclei). The aim is to obtain a reliable description for all nuclei deep into the regime of exotic ones. The predictive power of a parameterization is to be checked at three levels: 1) the ability to reproduce the fit-data, 2) the performance for interpolation to other nuclei in the range of the fit data, and 3) the reliability of extrapolations to other regions of the nuclear chart (e.g. super-heavy elements) or other observables. It is found that check 1 and check 2 are usually well satisfied while the extrapolation to super-heavy elements (SHE) reveals a systematic deviation. That holds for all modern Skyrme parameterizations. We will discuss {this issue} in terms of a newly developed fit protocol \cite{Klu09a}. Reference point is the Skyrme parameterization SV-min which was fitted to a large set of nuclei covering a wide span of mass numbers $A$ as well as long isotopic and isotonic chains. The fit pool selected good ``mean-field nuclei'', i.e. nuclei which have negligible effects from collective ground state correlations \cite{Klu08a}, and was confined to spherical systems for reasons of technical simplicity. The upper panel of figure \ref{fig:SV-min-def-energies} summarizes the error in binding energy for SV-min taken over all available nuclei. All energies are computed including collective ground state correlations. Filled squares indicate the fit nuclei (for which correlations are ignorable), open circles indicate well deformed nuclei (deformation $\beta_2>0.2$), and open triangles indicate the majority of vibrationally soft nuclei (vibrational amplitude larger than deformation, large correlations). The figure shows that interpolation (results for nuclei $A<210$) works nice with errors remaining acceptably small and distributed on both sides of the zero line. But the extrapolation to SHE shows a significant trend to increasing underbinding. The same trend (often worse) is found for other Skyrme forces. One could try to cure that defect by including data from SHE. This has been done by adding the energy of $^{264}$Hs (a well deformed SHE) to the fit data. This yields a modified parameterization ``SV-def'' whose distribution of errors on binding is shown in the lower panel of figure \ref{fig:SV-min-def-energies}. The predictions for other SHE (now being an interpolation) has clearly improved. But that is achieved at the price of sacrificing the quality of many other nuclei which are now often overbound. This indicates that there is an intrinsic problem with the form of the Skyrme energy functional which inhibits to span a wider mass range. \section{Fission barriers and half-lives of super-heavy elements} The microscopic description of nuclear fission is a long standing problem which was handled long ago in terms of empirical shell models, see e.g. \cite{Bra72aR}. The case is extremely demanding for self-consistent mean-field models as all aspects of the effective nuclear interaction are probed, global parameters of the nuclear liquid drop as well as details of the shell structure. SHF studies of fission are thus still rare, see e.g. \cite{Ber01a,Bur04,War06}. We have recently developed a fully self-consistent description of fission life-times \cite{Sch09a} and use it here to work out conflicting trends of fission properties in SHE. We summarize briefly the computational scheme as outlined in \cite{Sch09a}: The fission path is generated by quadrupole-constrained SHF whose energy expectation values yield a ``raw'' collective energy surface. The collective mass and moments of inertia are computed by self-consistent cranking along the states of the path \cite{Rei87aR}. Approximate projection onto angular momentum zero is performed using the moments of inertia and angular-momentum width. Quantum corrections for the spurious vibrational zero-point energy are applied (using quadrupole mass and width). The collective ground state energy is computed fully quantum mechanically \cite{Klu08a}. The tunneling rate at the given ground state energy and the repetition rates are computed by the standard semi-classical formula (known as WKB) using the quantum-corrected potential energy and collective mass; the fission life-time is finally composed from these two rates. All calculations are performed in axial symmetry. \begin{SCfigure}[0.7] { \epsfig{figure=figure4.eps, width=0.5\linewidth}} \caption{\label{fig:halflife-barrier3D} Fission barriers (lower) and half-lives (upper) for a selection of SHE as indicated. Experimental data are compared with results from a variety of different Skyrme parameterizations, SkM$^*$ \cite{Brack-skms}, SkP \cite{skp}, SkI3 \cite{ski3}, SV-bas and SV-min \cite{Klu09a}. The experimental data is taken from \cite{Hof01,Pet04,Oga04,Gre06,Gat08}. } \end{SCfigure} Figure \ref{fig:halflife-barrier3D} summarizes results on fission barriers and lifetimes for a few typical SHE and for a large variety of Skyrme parameterizations. The SHE represent two groups, one at the lower side and another one with much heavier nuclei at the limits of present days {available experimental} data. The span of predictions from the various Skyrme forces is huge in all cases in spite of the fact that all these parameterizations provide a high-level description of basic nuclear properties. But the variation of predictions is not the problem. One may decide to chose from the manifold of parameterizations just those which provide at the same time good fission properties throughout. But this turns out to be impossible at present. The true problem becomes apparent when looking at the trend from the lighter side (Rf, Sg, Hs) to the heavier elements (Z=112, 114). All parameterizations produce a wrong trend of the predictions from the lower to the upper region. Forces which perform acceptable for Rf, Sg, Hs fail badly for Z=112,114 and vices versa. One may argue that triaxiality, ignored here, could resolve the trend because triaxial deformation may lower some barriers selectively. But that is very unlikely in view of the experience that the triaxial barrier-lowering amounts typically to 1 MeV, at most 2 MeV \cite{Cwi96a}, which does not suffice to bridge the gap here. \section{Conclusion} We have worked out briefly five puzzling points in connection with SHF:\\ 1) An interpretation as ``force'' is inconsistent because the density dependence inhibits an expression of the Skyrme energy as standard quantum-mechanical expectation value.\\ 2) The center-of-mass correction, usually applied, causes conceptual problems in nuclear fusion, fission and fragmentation; parameterizations which are used for such reactions should be adjusted without including the center-of-mass correction.\\ 3) It is presently impossible to find a parameterization which delivers a good description of the giant dipole resonance in all regions of the nuclear chart; it seems that the relation of surface to volume mode is not properly balanced.\\ 4) The extrapolation of binding energies to SHE yields quickly increasing underbinding and a refit including energies of known SHE spoils the quality in the region of stable nuclei; the problem is probably caused by a still inappropriate surface or curvature energy.\\ 5) The experimentally observed trend of fission properties from the Hs region to much heavier SHE is not reproduced by any SHF parameterization.\\ For all these points, we do not have presently any solution and often we have not even figured {out} the deeper reasons. This has to be put on the work schedule for future studies. \bigskip \noindent \section*{Acknowledgment} We thank the regional computing center of the university Erlangen-N\"urnberg for generous supply of computer time for the demanding calculations. The work was supported by the BMBF under contracts 06 ER 808 and 06 ER 9063. \bigskip \section*{References}